You are on page 1of 34

Journal of Physics B: Atomic, Molecular and Optical Physics

TOPICAL REVIEW • OPEN ACCESS Related content


- Attosecond physics at the nanoscale
Attosecond physics phenomena at nanometric tips M F Ciappina, J A Pérez-Hernández, A S
Landsman et al.

To cite this article: Michael Krüger et al 2018 J. Phys. B: At. Mol. Opt. Phys. 51 172001 - Attosecond physics in photoemission from
a metal nanotip
M Krüger, M Schenk, M Förster et al.

- Interaction of ultrashort laser pulses with


metal nanotips: a model system for strong-
View the article online for updates and enhancements. field phenomena
Michael Krüger, Markus Schenk, Peter
Hommelhoff et al.

This content was downloaded from IP address 125.161.104.138 on 20/02/2019 at 12:43


Journal of Physics B: Atomic, Molecular and Optical Physics

J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 (33pp) https://doi.org/10.1088/1361-6455/aac6ac

Topical Review

Attosecond physics phenomena at


nanometric tips
Michael Krüger1,5 , Christoph Lemell2 , Georg Wachter2,
Joachim Burgdörfer2 and Peter Hommelhoff3,4
1
Department of Physics of Complex Systems, Weizmann Institute of Science, 234 Herzl Street, Rehovot
76100, Israel
2
Institute for Theoretical Physics, Wiedner Hauptstraße 8-10, Vienna University of Technology, 1040
Vienna, Austria
3
Department of Physics, Friedrich-Alexander-Universität Erlangen-Nürnberg (FAU), Staudtstraße 1,
91058 Erlangen, Germany
4
Max-Planck-Institut für die Physik des Lichts, Staudtstraße 2, 91058 Erlangen, Germany

E-mail: michael.krueger@weizmann.ac.il, lemell@concord.itp.tuwien.ac.at and peter.hommelhoff@


physik.uni-erlangen.de

Received 16 February 2018, revised 15 April 2018


Accepted for publication 22 May 2018
Published 9 August 2018

Abstract
Attosecond science is based on electron dynamics driven by a strong optical electric field and has
evolved beyond its original scope in gas-phase atomic and molecular physics to solid-state targets. In
this review, we discuss a nanoscale attosecond physics laboratory that has enabled the first
observations of strong-field-driven photoemission and recollision at a solid surface: laser-triggered
metallic nanotips. In addition to the research questions of rather fundamental nature, femtosecond
electron sources with outstanding beam qualities have resulted from this research, which has prompted
follow-up application in the sensing of electric fields and lightwave electronics, ultrafast microscopy
and diffraction, and fundamental matter-wave quantum optics. We review the theoretical and
experimental concepts underlying near-field enhancement, photoemission regimes and electron
acceleration mechanisms. Nanotips add new degrees of freedom to well known strong-field
phenomena from atomic physics. For example, they enable the realization of a true sub-optical-cycle
acceleration regime where recollision is suppressed. We also discuss the possibility of high-harmonic
generation due to laser irradiation of metallic nanostructures.

Keywords: Attosecond physics, nano-optics, strong-field physics


(Some figures may appear in colour only in the online journal)

1. Introduction and early developments Ferray et al 1988) has led to the invention and broad accep-
tance of the recollision picture, which is at the core of the so-
The discovery of the surprisingly efficient generation of high- called simple man’s model or three-step model. While its
harmonic radiation from atomic gases (McPherson et al 1987, origins date back about 30 years (Kuchiev 1987, Galla-
gher 1988, van Linden and van den Heuvell 1988), the model
5
Author to whom any correspondence should be addressed. was widely accepted about 25 years ago when a number of
papers appeared (Corkum 1993, Schafer et al 1993, Becker
et al 1995). It is thus appropriate to celebrate the 25th anni-
Original content from this work may be used under the terms
versary of the recollision picture with this special issue. Since
of the Creative Commons Attribution 3.0 licence. Any
further distribution of this work must maintain attribution to the author(s) and then, several new and quite spectacular research fields have
the title of the work, journal citation and DOI. grown out of the initial discoveries, which originated in the

0953-4075/18/172001+33$33.00 1 © 2018 IOP Publishing Ltd Printed in the UK


J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

realm of high-field atomic physics but have now reached out particular in cold field emission mode, is extremely bright and
into various other disciplines (see Calegari et al (2016) for an coherent, which is the reason it is routinely used in high-
overview of recent developments). Examples include the resolution electron microscopes, even though the easier-to-
efficient generation of coherent extreme ultraviolet (XUV) operate Schottky sources are nowadays preferred despite their
and even x-ray pulses, the generation of attosecond light slightly inferior performance (Spence 2013). A highly
pulses, electron diffraction experiments with the electrons coherent electron beam can be generated in DC operation
emitted from the irradiated atoms, molecules and clusters, and when a strong static negative electric field is applied to a
strong-field and recollision physics in condensed matter. In metal tip. Because of the accumulation of field lines at the tip
this contribution, we provide a review on this last topic, apex, the potential landscape in the vicinity of the apex is
namely strong-field and recollision physics at the surface of changed so that electrons can tunnel through the potential
nanoscale solids. Aspects of this research are reflected in a barrier from the metal to the vacuum side. This is the well
number of reviews and books (Vasa et al 2009, Krüger known field emission process, requiring static field strengths
et al 2012a, Hommelhoff and Kling 2015, Jones et al 2016, in excess of 1 GV m−1 to observe a measurable electron
Ciappina et al 2017). The focus of our review will be on a current from a typical metal.
particular nanostructure geometry: sharp metallic needle tips. Nonlinear photoemission, in particular the two con-
We will recapitulate how three well-established fields ceptually identical processes of above-threshold ionization of
came together to merge into a new field, offering new insight atoms and above-threshold photoemission from solids,
and applications alike: (1) a century of experience in field represents the foundation on which strong-field physics
emission from needle tips, (2) atomic recollision physics, and stands. We recall early work on laser-driven (multi-) photon–
(3) nano-optics, also known as near-field optics. This com- electron emission from solids, before we go in medias res.
bination enables an unprecedented confinement of light– With ‘early’ we refer to work published before the three
matter interaction to nanometer length and attosecond time 2006/2007 papers, which can be considered the first report on
scales. coherent electron sources driven by ultrafast photoemission
processes (Hommelhoff et al 2006b, 2006c, Ropers et al
1.1. Overview and structure 2007c). Subsequent papers provided clear evidence of strong-
field physics at metal tips (Bormann et al 2010, Schenk
In the following part of the introduction (section 1.2), we will
et al 2010).
review early work on needle tips and strong-field photo-
Multiphoton photoemission from metals dates back to the
emission from solids. Because nano-optics is of high impor-
seminal work by Keldysh in 1965 (Keldysh 1965). In the
tance for strong-field physics at needle tips, we will discuss it
same year, Keldysh’s method was applied to metals by
in greater detail in Chapter 2. Chapter 3 deals with exper-
Bunkin and Fedorov, who, in essence, showed that the lim-
imental methods that have been used to elucidate the many
iting cases are DC field emission in the low-frequency limit
open questions around the merger of gas-phase strong-field
physics with solid-state systems. In Chapter 4, we discuss and multiphoton physics for high frequencies (Bunkin and
photoemission mechanisms, the in-depth understanding of Fedorov 1965), closely mirroring analogous concepts in
which is not only important in its own right but for realizing atomic physics.
new ultrafast and coherent electron sources. In Chapter 5, an On the experimental side, the invention of the laser
overview is given of theoretical methods that are frequently quickly led to the discovery of nonlinear photoemission; two-
used for the modeling of strong-field physics at metal tips. photon photoemission from solids was first observed in 1964
Temporal dynamics of the emission processes is discussed in at a semiconducting Cs3Sb photocathode (Sonnenberg
more detail in Chapter 6, while prototypical strong-field et al 1964). In Budapest, Farkas and colleagues investigated
effects, namely the acceleration and recollision of photo- higher-order nonlinear photoemission early on with pico-
electrons by the waveform of the driver field, are reviewed in seond laser pulses (Farkas et al 1971, 1972, Farkas and
Chapter 7. Modulating the waveform enables the control of Chin 1985), referring to Keldysh’s theoretical work. Clear
strong-field photoemission and acceleration, as discussed in deviations from perturbative multiphoton emission and the
Chapter 8. Chapter 9 explores the feasibility of high-harmonic onset of optical tunneling effects in electron emission from a
generation (HHG) at nanotips and presents the first ab initio gold surface were reported in 1991 (Tóth et al 1991).
results. In Chapter 10, we give a brief overview of the wide In parallel, the photofield emission process was investi-
range of applications of the observed strong-field phenomena gated and later utilized for inferring the band structure of metals.
at needle tips, before concluding our review in Chapter 11. Here, an electron is photo-excited inside the metal close to or
below the emission threshold, and then undergoes field emis-
sion (Lee 1973, Radoń 1998). Because field emission is
1.2. Early work
exponentially sensitive to the barrier height, a current mea-
Sharp metal needle tips have already been used for more than surement as a function of the applied DC voltage allows the
80 years, even before the pioneering works of Erwin W identification of the origin of the electrons; this, in combination
Müller (Müller 1936), as excellent electron sources in various with the known but variable photon energy, enables a recon-
applications. The electron beam emitted from such a tip, in struction of the band structure of the metal under scrutiny.

2
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

While the well-defined excitation of electrons (from close over thermionic emission was made possible by excitation
to the Fermi level) to a well-defined intermediate state is the with laser pulses with a duration below 200 fs (Ferrini
basis of photofield emission spectroscopy, the (laser-pulse- et al 2009).
induced) heating of the electron gas inside a metal tip also Around the year 2000, the invention of the frequency
leads to enhanced electron emission rates (Lee et al 1980, comb allowed control of the optical carrier field within the
Riffe et al 1993). Hence, spectral information of the emitted laser-pulse envelope (Udem et al 1999, Diddams et al 2000,
electrons is mandatory to obtain a clear picture of the emis- Jones et al 2000, Udem et al 2002, Hänsch 2006). This was
sion process and its dynamics. In particular, the question enabled by controlling the relative phase between the carrier
arises whether the participating electrons thermalize before field maximum and pulse envelope maximum, the carrier-
they are emitted, or whether the electrons are emitted coher- envelope offset phase, or, in short, carrier-envelope phase
ently as indicated by photon orders showing up in the spectra. (CEP). In the strong-field regime with atoms, it allowed full
However, an unstructured spectrum does not necessarily control over the electron dynamics on attosecond time scales
imply that the emission is thermal and incoherent, as photo- and the generation of isolated attosecond XUV light pulses
emission studies from flat metals are prone to focal averaging (Baltuška et al 2003, Kienberger et al 2004, Sansone
effects. Focal averaging implies that electrons emitted in et al 2006). Lemell et al (2003) performed a numerical
various regions of the laser spot experience different peak investigation of the CEP dependence of the electron emission
laser field strengths leading to a smeared-out electron energy current from a metal surface based on a jellium model. They
distribution, to the extent that photon orders are no longer predicted a significant CEP-dependent current for various
discernible. Yet another effect often arising in photoemission intensity regimes, in particular for low intensities in the
in the high-intensity regime from flat metal surfaces can wash intermediate regime between multiphoton and tunneling
out photon orders, namely space-charge repulsion within the photoemission. Apolonski et al (2004) indeed observed a
emitted electron cloud. The reader is referred to earlier work CEP-dependent current contribution from a flat metal surface
on flat metal surfaces, for example, to Aeschlimann et al illuminated with sub-two cycle laser pulses, but it was found
(1995), Damascelli et al (1996), and to Gault et al (2007), to be very small. In theoretical studies, Faisal et al (2005)
Kealhofer et al (2012) for more recent work on metal nee- predicted electron emission spectra from solid surfaces that
dle tips. bear reminiscence of the hallmark high-harmonic plateau well
Two-photon photoemission (2PPE), in particular time- known from atomic physics. They indeed relate their num-
resolved 2PPE, is the method of choice for inferring various erical observations to the recollision plateau in atomic physics
(often electronic) time scales in metals. It has strong ties to —the first study to note that recollision of electrons might
strong-field physics at metal surfaces, given the fact that this appear at a metal surface.
method allows the measurement of electronic processes on Around the same time spectrally resolved experiments
the femtosecond scale. We will not discuss this broad research were performed on electron emission from flat metal surfaces
field further, but refer the reader to two reviews (Petek and with femtosecond laser pulses, revealing above-threshold
Ogawa 1997, Bauer et al 2015). photon orders, i.e. electrons emitted with kinetic energies
Laser pulses with sub-picosecond duration are necessary larger than at least the energy of a driving-laser photon (Banfi
for these studies to reach the nonlinear photoemission regime et al 2005, Bisio et al 2006). The observation of above-
without inducing damage to the sample. The large deviation threshold orders represents the first step towards observing
of the electron temperature from the lattice temperature on the exciting physics involving electrons with energies much
sub-picosecond time scales is indeed important for explaining higher than needed just for their emission from the solid,
the observed emission behavior in photoemission experiments including the tell-tale recollision plateau. Signatures of non-
with sub-picosecond pulses (Riffe et al 1993, Girardeau- linear photoemission, though not spectrally resolved, were
Montaut and Girardeau-Montaut 1995). found in the initial work on femtosecond laser-driven emis-
Long before the recollision picture had been introduced, sion from nanometer sharp needle tips performed in Stanford
above-threshold ionization had been observed from atoms (Hommelhoff et al 2006b, 2006c) and Berlin (Ropers
(Agostini et al 1979, Kruit et al 1981), where many photon et al 2007a, 2007c). Based on the previous work discussed
orders could be unequivocally identified in photoionization above, it was clear that a strong CEP-dependent current from
spectra. Similarly, shifting of the multiphoton peaks because metal surfaces, in particular from needle tips, should be
of field effects, in particular the AC Stark effect and the role observable, and indeed initial simulations hinted in that
of the ponderomotive potential, had been well understood direction (Hommelhoff et al 2006a, 2007, Stockman and
through experiments from atoms in the gas phase (Kruit Hewageegana 2007).
et al 1983, Muller et al 1983, Bucksbaum et al 1987). Some We note that (pulsed) electron sources are often
ten years later, indications of the counterparts in photoemis- responsible for limiting the maximum achievable beam
sion from solids had been reported: above-threshold photo- brightness in accelerator-related applications. For this reason,
emission (ATP) including potential ponderomotive effects much research has been devoted to pulsed low-emittance
initially in the late 80s (Luan et al 1989, Fann et al 1991, electron sources, but this is outside of the scope of this
Farkas et al 1993) and later more clearly (Banfi et al 2005, review. We refer the reader to the original work based on
Bisio et al 2006). Notably, well-defined photon orders could needle sources (Boussoukaya et al 1989, Hernandez Garcia
be observed, and the prevalence of prompt coherent emission and Brau 2002, Ganter et al 2008). Similarly, in atom probe

3
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

tomography, various materials and alloys are cast into needle


form. With either (positive) high-voltage pulses or laser pul-
ses, the tip is decomposed atom by atom. The fragment atoms
are individually analyzed to reconstruct the exact morphology
of the parent material, which is of utmost interest to material
science. Laser-pulse-induced heating of needle tips is there-
fore also investigated in this field (see Miller et al 1996,
Vurpillot et al 2009, Gault et al 2012).

2. Near-field optics at nanotips

Nanometric needle tips give rise to optical near-fields on the


Figure 1. Nano-optical near-fields at nanostructures. (a) Illustration
nanoscale, strongly enhancing the electromagnetic field at of a near-field at a metallic nanosphere induced by a laser field (red).
sharp geometrical features. Field enhancement plays a key (b) Normalized local electric field strength at a gold nanosphere,
role in opening up strong-field physics to solid-state systems λ=720 nm.
for three reasons. First, the requirements on the pulse energy
of the laser source are significantly lowered, so that an
oscillator system with as little as 100 pJ pulse energy is suf- optics, the electromagnetic response of the material to exter-
ficient to observe strong-field photoemission. Second, the nal fields is given by the complex dielectric function
reduced pulse energy and the tiny geometrical cross-section  (l ) =  r (l ) + i i (l ). ò is the square of the complex
of the nanotips help to avoid excessive heat deposition and refractive index. Assuming a homogeneous static electric field
thermal damage when irradiated with strong laser pulses. By around the nanosphere, a collective displacement of electric
contrast, strong-field experiments at extended solid surfaces charge is induced inside the material with respect to the ionic
are challenging due to focal averaging, thermal effects and background. In alternating electric fields, that is, in optical
space-charge broadening. Third, the optical near-field loca- fields, this displacement starts to oscillate, leading to a time-
lizes photoemission to sub-micron emission areas, enabling dependent polarization (see figure 1(a)). The sphere then acts
measurements of photocurrents from a single well-defined as a nano-emitter of light.
emitter. In the following, we will briefly introduce the reader For a quantitative description of the near-field, it is
to the main concepts of near-field optics, before focusing in instructive to neglect its time dependence and weak magnetic
depth on nanotips. field components and make use of the quasi-static approx-
imation (Jackson 1999, Maier 2007). Figure 1(b) shows the
calculated near-field at a gold nanosphere with a radius of
2.1. Introduction to near-field optics
30 nm when applying an optical field with λ=720 nm.
For many applications such as microscopy and imaging, it is Three main observations can be made. First, at the surface of
of considerable interest to confine the light–matter interaction the sphere the electric field is strongly enhanced. Second, at
to small volumes and length scales. Two length scales are increasing distances from the surface into free space, the
involved—the optical wavelength and the dimensions of the magnitude of the field rapidly decays and eventually returns
object. The first length scale is bound by Abbe’s diffraction to that of the externally applied field. Third, inside the sphere
limit of a focused light beam. The second length scale—once the local field is homogeneous and its magnitude is much
the object’s geometric features are chosen to be much smaller smaller than that of the applied field—a screening effect.
than the optical wavelength—leads to strong confinement of These three effects, field enhancement, field localization and
the optical field, enabling optics far below the diffraction screening result from charge redistribution and together form
limit. Electromagnetic near-fields are induced that are loca- the pillar of nano-optics.
lized near the object’s surface. Nano-optics, also called near- The maximum local field enhancement ξ, defined as the
field optics, is based on this property and is an active research magnitude ratio of the total field relative to the applied field,
field in itself (see, e.g. Maier 2007, Sarid and Challener 2010, is found at the poles of the sphere and is given by
Novotny and Hecht 2012). Nano-optics enables a range of  (l ) - 1
applications in microscopy and spectroscopy, for example x= 1+2 . (1 )
 (l ) + 2
scanning near-field microscopy (see Wessel (1985), Inouye
and Kawata (1994), Hartschuh (2008)) and tip-enhanced Most significantly, ξ is independent of the radius of the sphere
Raman scattering (see Wessel (1985), Stöckle et al (2000)). for spheres that are small compared to λ, but strongly
In order to illustrate near-field enhancement, we focus on dependent on the dielectric properties of the material and
the example of a nanosphere as the simplest of all nanometric wavelength. The enhancement factor for the gold nanosphere
geometries. The nanosphere is made of an isotropic material at its optical ‘hotspots’ is ξ=3.41 at 720 nm, corresponding
and is exposed to monochromatic light of wavelength λ in to a local intensity enhancement of ξ2=11.6. Nonlinear
vacuum. The nanosphere’s radius R is much smaller than λ, processes are strongly enhanced at these hotspots, enabling
producing sub-wavelength spatial confinement. In linear low-order harmonic generation (Bouhelier et al 2003, Neacsu

4
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

et al 2005a, Wolf et al 2016) and strong-field photoemission field enhancement on both the tip radius and opening angle.
(Bormann et al 2010, Schenk et al 2010, Dombi et al 2013, Unlike for nanospheres (see equation (1)) and ellipsoids there
Keathley et al 2013) for moderately energetic laser pulses or are no analytical formulae available for the field enhancement
even continuous wave lasers (Sivis et al 2018). at the nanotip geometry. The nanostructure closest to a
Nanostructures exist in a large variety of shapes, much nanotip for an analytical solution within the quasi-static
more complicated than spheres. Frequently used examples are approximation is an ellipsoid (Novotny and Hecht 2012). One
nanotips, nanorods, nanotriangles and composite structures therefore relies on numerical methods to solve Maxwell’s
such as bow-tie antennae and dense nanoarrays. Classical equations for a realistic incident light field incorporating the
electrodynamics according to the linear Maxwell equations is, nanotip’s geometrical boundary conditions. Widely used
in most cases, sufficient to understand and simulate the nano- numerical methods in nano-optics are the finite element
optical response of such structures. Generally, near-field method, the finite-difference time-domain (FDTD) method or
formation critically depends on the polarization and wave- the boundary element method (BEM) (Taflove and Hagness
length of the incident light, the geometry of the nanostructure 2005). In the context of strong-field photoemission, both
and on the (wavelength-dependent) dielectric constant of the FDTD and BEM have been successfully applied (see, for
material ò(λ). The magnitude of the field enhancement is example, the works of Yanagisawa et al (2009, 2010), Thomas
related to the absolute value of the complex dielectric func- et al (2013, 2015), Förg et al (2016), Ahn et al (2016, 2017)
tion: the larger ∣ ∣, the larger ξ becomes. for FDTD and the study of (Thomas et al 2015) for BEM). In
The appearance of nano-optical hotspots is due to the addition, a discontinuous Galerkin time-domain method has
geometry defining the boundary conditions to Maxwell’s
been developed for nanotips (Swanwick et al 2014). Finite
equations. The external light field induces charges at dis-
element methods have been used for simulating static fields
continuities at material interfaces and boundaries. If the polar-
around a nanotip once it is integrated into an electron optical
ization of the external field is aligned with the normal direction
device (see, e.g. Paarmann et al 2012, Hoffrogge et al 2014,
of sharp features, these charges give rise to strong local fields.
Bormann et al 2015 McNeur et al 2016, Storeck et al 2017).
The smaller the length scale of those sharp features such as their
Figure 2(a) shows the calculated field strength as a
local radius of curvature with respect to the incident wave-
function of position for a tungsten tip with R=10 nm and
length, the larger the field enhancement becomes. This mech-
θ=15° and an incident laser pulse with a full width at half
anism is the time-dependent analog of the electrostatic
maximum (FWHM) intensity duration of τ=5 fs and central
lightning-rod effect and is a common feature for all materials.
wavelength λ=800 nm (Thomas et al 2015). The field is
A second geometric effect is strongly related to the wave-
length of the light. A linear extension of an odd multiple of half polarized parallel to the tip axis. In this configuration, the
of the wavelength can cause field enhancement significantly main features of near-field optics, namely field enhancement,
surpassing the other mechanisms creating an antenna resonance. localization and screening of the incident optical field become
Nano-plasmonics adds a third mechanism to near-field apparent. The tip is acting as a highly effective antenna for the
formation, namely localized surface plasmons (LSPs). Refer- optical field and exhibits strong-field enhancement, with
ring to the simple example of the nanosphere, equation (1) ξ∼6.5, close to the apex of the tip. Away from the tip
predicts a resonance at òr=−2, a condition called Fröhlich surface, along the symmetry axis, the near-field strongly
condition (Bohren and Huffman 1998). The Fröhlich condition decays near-exponentially and relaxes to the field strength of
is met by plasmonic metals in the visible spectral region the incident field. The decay length lF is typically of the order
(òr<0 and 0 <  i  ∣ r∣), such as gold and silver. LSPs can lF∼0.8 K 0.9R and scales approximately linearly with R.
be excited by direct irradiation or by coupling light through a The value of lF is crucial when describing sub-optical-cycle
grating imprinted on the nanostructure (see, e.g. Ropers electron acceleration driven by the near-field (see Chapter 7).
et al 2007b, Berweger et al 2012). Applying the latter method Inside the tip, screening results in a highly reduced field.
to a plasmonic nanotip leads to a traveling surface plasmon, Nanotips for strong-field photoemission can be fabricated
which is focused and ultimately localized at the tip’s apex from various conductive materials, such as tungsten, gold,
(Stockman 2004). With the light on resonance, this naturally silver, aluminum or doped silicon (see Chapter 3 for methods
leads to higher field enhancement than other mechanisms of tip fabrication and characterization). As discussed above, it
operating in non-plasmonic materials. However, as for every is desirable to either choose a dielectric material with a large
resonance phenomenon, the build-up of the resonance requires absolute value of the dielectric function ò(λ) or to employ an
a characteristic time, which may limit the suitability of such LSP resonance in a plasmonic material. Figure 2(b) displays
resonant structures for ultrafast processes. the wavelength-dependent dielectric function of various mate-
rials. For tungsten, gold and aluminum, moving to longer
2.2. Near-fields at nanotips
wavelength in the near-infrared or mid-infrared domain will
Nanotips are tapered needles that end in an approximately increase ∣ ∣ and hence the field enhancement. The increase is
conical apex. Usually they are modeled as a hemispherical most pronounced for aluminum. Already at 800 nm, the di-
apex with radius of curvature R attached to a tapered shank electric function is ò=−64+47i. In addition, gold provides
with half-opening angle θ (see figure 2(a) for an illustration). plasmonic near-field enhancement because its dielectric tra-
The presence of the tip shank introduces a dependence of the jectory in the complex plane (figure 2) passes near òr=−2, the

5
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Figure 2. Near-fields at nanotips. (a) An FDTD calculation solving Maxwell’s equations yields the local electric field strength at the time when the
near-field induced by a 5 fs, 800 nm laser pulse is at its maximum. The nanotip is characterized by the tip radius R and the (half-)opening angle α (here
R=10 nm and θ=15°). The laser field E is linearly polarized along the tip’s pointing direction and travels in the y direction. Gray arrows indicate
the local field direction. (b) Dielectric function for various materials. Real parts òr and imaginary parts òi of the complex dielectric function for tungsten
(W), aluminum (Al), gold (Au) and silicon (Si) as functions of wavelength λ (color). (c) Maximum field enhancement ξ for tungsten and gold nanotips
with varying radius and opening angle (τ=5 fs, λ=800 nm). Adapted from Thomas et al (2015). © IOP Publishing Ltd. CC BY 3.0.

Fröhlich resonance condition for nanospheres (equation (1)).


Note that doped semiconductors such as silicon can also
exhibit strong-field enhancement.
Figure 2(c) shows the geometry dependence of the
maximum field enhancement as a function of R and θ for
tungsten and gold tips irradiated by a typical ultrashort laser
pulse of 5 fs duration and λ=800 nm (Thomas et al 2015).
Three main observations can be made from the numerical
FDTD calculation. First, ξ decreases monotonically with
increasing R for both materials and for any given θ, as
expected. The dependence on R is confirmed by measure-
ments performed on tungsten and gold tips (Thomas
et al 2013, Krüger et al 2014; see figure 3). Second, θ has a
strong influence on the field enhancement, with maxima
around θ=40° and θ=15° for tungsten and gold, respec-
tively. The non-monotonic dependence on θ and the appear-
ance of a maximum at intermediate θ for both dielectric and
plasmonic materials can be traced to the build-up of the
Figure 3. Measured field enhancement factors at tungsten and gold
surface charge density. Its magnitude is controlled by the nanotips. Field enhancement factor of tungsten tips (blue dots) and
interplay between the illuminated surface area increasing with gold tips (red squares) as a function of the tip radius. Uncertainty in ξ
θ and the normal component of the electric field decreasing represents an estimated systematic error due to the uncertainty in
with θ, as predicted by electrostatics. For plasmonic materials, laser intensity. Lines show the results of a numerical calculation of
this effect is further enhanced by the excitation of a surface Maxwell’s equations for 800 nm 5.5 fs laser pulses (W: solid blue
line, Au: dashed red line). Reproduced from Krüger et al (2014).
plasmon, where the resonance condition linking the dielectric © IOP Publishing Ltd. All rights reserved.
constant, wavelength and opening angle leads to a maximum
at smaller angles. This enhancement favors the fabrication of than for tungsten with ξ=12 at θ=35° due to the plas-
tips with small radii and with opening angles near the max- monic response.
imum. Third, the maximum enhancement in the calculated The collective electronic response of the medium leads to
parameter space is ξ=36 at θ=15° for gold, much larger a phase shift Δf between the incident field and the resulting

6
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

applied to measure the amplitude and phase of a near-field at a


nanotip (Förg et al 2016). Recently, an experiment using
terahertz streaking successfully resolved surface plasmon
propagation at a nanotip (Wimmer et al 2017).
So far, we have discussed optical near-fields within the
scope of classical electrodynamics, which is sufficient for
most applications. However, quantum effects can become
important when considering nanostructures, opening up
‘quantum plasmonics’ (see Zhu et al (2016) for a recent
review). For example, material-vacuum boundaries should
not be considered to be infinitely thin. Instead, the electron
density at the surfaces is smeared out and slightly leaks into
the vacuum (Zuloaga et al 2010). This can lead to an effective
decrease of the magnitude of field enhancement (Zuloaga
et al 2010, Ciracì et al 2012) and can affect plasmonic
Figure 4. Typical experimental setup for strong-field photoemission antenna resonances in the nano-gap between two nanos-
from a metal nanotip. A linearly polarized beam consisting of tructures (Marinica et al 2012, Savage et al 2012, Scholl
femtosecond laser pulses (red) is focused on the apex of a metallic et al 2013, Marinica et al 2015). Ab initio theory approaches,
nanotip with the help of an off-axis parabolic mirror (OAP). such as time-dependent density functional theory (TDDFT),
Electrons (blue) are emitted in forward direction and are detected are able to provide a self-consistent treatment including such
either using a micro-channel plate (MCP) detector for spatial
resolution or an electron spectrometer for spectral resolution. The quantum effects, but suffer from technical limits of current
transmitted laser beam can be used for diagnostics. computing power. An alternative approach introduces
quantum modifications to Maxwell’s equations (Esteban
optical near-field. The magnitude of this retardation effect et al 2012, 2015). To the best of our knowledge, these effects
also depends on the nanotip geometry and material, but not as have not been observed for strong-field-driven nanostructures.
strongly as ξ. For the parameters used in figure 2(c), the The measurement shown in figure 3 is consistent with the
simulations show that tungsten and gold exhibit shifts of description by Maxwell’s equations.
Δf∼0.3 ... 0.6π and Δf∼0.4 ... 0.7π, respectively. Under Experiments towards and in the strong-field regime were
pulsed irradiation, the optical response of the nanotip also carried out mostly with nanotips, but there are related
leads to small changes in the central wavelength and pulse experiments with other types of nanostructures, such as free-
standing nanowire tips with sharp edges (Ahn et al 2017),
duration. In the case of a plasmonic response, long-lived
nanospheres (Schertz et al 2012), nanorods (Sun et al 2013,
plasmon oscillations might persist after excitation (see, e.g.
Kusa et al 2015, Lehr et al 2017), nanotriangles (Putnam
Sönnichsen et al 2002). Nonlinear effects such as second
et al 2017), nanostars (Sivis et al 2018), carbon nanotubes
harmonic generation (Bouhelier et al 2003) can also influence
(CNTs) (Li et al 2017), composite bow-tie antennas and
the optical near-field.
nanorod antennas (Dombi et al 2013, Rybka et al 2016,
Measuring the strength and shape of near-fields at
Hobbs et al 2017, Putnam et al 2017, Rácz et al 2017). In
nanotips and other nanostructures is possible using a variety
most of these experiments, the nanostructures were located on
of methods. Nonlinear processes driven by the near-field such
substrates and were made from gold, hence enabling plas-
as second harmonic generation (Bouhelier et al 2003, Neacsu monic field enhancement when on resonance. In addition,
et al 2005a), two-photon photoemission (Tsujino et al 2009), gas-like ensembles of dielectric nanospheres (Zherebtsov
multiphoton photoemission and strong-field photoemission et al 2011, Süßmann et al 2015, Seiffert et al 2017a, 2017b),
(Hommelhoff et al 2006b, Ropers et al 2007c, Grubisic metallic nanoclusters (Passig et al 2017) and C60 ‘buckyballs’
et al 2013, Park et al 2013, Thomas et al 2013, Krüger (Li et al 2015) were used in strong-field experiments.
et al 2014, Rácz et al 2017) are extremely sensitive to the
enhancement. Some of the latter investigations also exploit
electron acceleration or recollision driven by the near-field in 3. Experimental methods
order to retrieve the local intensity (for more details, see
Chapter 10). Other methods rely on the interaction of a tightly In this chapter, we briefly introduce the experimental methods
focused high-energy electron beam passing by the nanos- that enable strong-field studies at nanotips and other nanos-
tructure. The near-field manifests itself in triggering cath- tructures. Figure 4 shows a typical experimental setup.
odoluminescence (Vesseur et al 2007, Chaturvedi et al 2009), Femtosecond laser pulses are focused onto the apex of a
incoherent energy loss of the electron beam (Nelayah metallic nanotip situated in an ultrahigh vacuum chamber.
et al 2007, Huth et al 2013, Schröder et al 2015b, Yalunin Pressures ideally below 10−8 Pa are helpful to maintain
et al 2016) or coherent electron energy loss and gain (Bar- atomically clean surfaces over at least one hour. The laser
wick et al 2009, Feist et al 2015), enabling its spatio-temporal polarization axis is parallel to the tip’s symmetry axis,
characterization. Attosecond nanoplasmonic streaking enabling maximum field enhancement and pronounced pho-
(Stockman et al 2007, Süßmann and Kling 2011), has been toemission in the forward direction. Photoelectrons are

7
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

usually detected using either a detector with spatial resolution investigations employed molybdenum nanotip arrays (Mus-
to image the emission pattern or an electron spectrometer to tonen et al 2011), gold nanorod arrays (Dombi et al 2013),
measure the photoelectron spectrum. In the following, we will pillar arrays (Nagel et al 2013) and junction devices (Rybka
focus on each element in more detail. et al 2016), silver tips (Bionta et al 2016), doped silicon tip
Initial experiments used widely available titanium sap- arrays (Swanwick et al 2014), hafnium carbide nanotips
phire-based laser sources at a center wavelength around (Kealhofer et al 2012) and carbon cone nanotips (Bionta
800 nm generated by both high repetition rate laser oscillator et al 2014). In addition, carbon nanotube (CNT) arrays were
and low-repetition rate amplifier systems. In oscillator-based used for strong-field experiments (Li et al 2017).
experiments, the damage threshold of the tungsten nanotip Detection is typically focused either on spatial resolution
prevented the use of enhanced peak intensities Ieff of more or spectral resolution. Spatial information can be gained in a
than 2 ´ 1013 W cm-2 (Krüger et al 2011). Using kHz field emission microscope setup where the nanotip is com-
repetition rate systems, higher intensities could be reached, bined with an MCP detector with a phosphorus screen,
for instance Ieff ~ 5 ´ 1013 W cm-2 for 30 fs, 800 nm pulses located a few cm away from the tip (see figure 4 and Yana-
(Bormann et al 2010) and a similar value for 4 fs, 700 nm gisawa et al (2009) for example experimental setups). The
pulses (Hoff et al 2017b). Working towards applications of spatial emission pattern of field emission or photoemission is
laser-triggered nanotips as coherent electron guns, blue or UV mapped on the MCP with a spatial magnification of up to 106,
laser pulses were employed (see, for example, Ehberger enabling nanometric resolution without additional electron
et al 2015, Bormann et al 2015, Storeck et al 2017). On the optics. Chevron-type MCPs allow detecting emitted electrons
high-frequency frontier, single isolated attosecond XUV with a typical efficiency of roughly 50%. Spectral resolution
pulses were applied in order to perform attosecond nano- is achieved by electron spectrometers either based on time-of-
plasmonic streaking (Förg et al 2016). In the mid-infrared flight measurements of laser-triggered electrons (Hilbert
spectral region, experimental studies relied on fiber lasers et al 2007, 2009, Keathley et al 2013, Hoff et al 2017b) or on
(see, e.g. Thomas et al 2012, Rybka et al 2016, Putnam electrostatic techniques. Typically, retarding field (see, e.g.
et al 2017) or tunable systems based on optical parametric Schenk et al 2010, Herink et al 2012) or hemispherical
amplification (see, e.g. Herink et al 2012, Park et al 2012, analyzers (Yanagisawa et al 2011, Park et al 2012, 2013,
Homann et al 2012, Piglosiewicz et al 2014, Förster Piglosiewicz et al 2014, Yanagisawa et al 2016) are used.
et al 2016, Schötz et al 2018). Also, THz pulses were added Hemispherical electron analyzers have an additional advan-
to the repertoire of light sources, effectively providing a quasi tage as they also provide angular resolution. In combination
DC electric field transient strongly localized at the nanotip with an imaging screen around their entrance aperture, they
(Herink et al 2014, Wimmer et al 2014, Li and Jones 2016, furthermore allow for site-selective spectral measurements.
Wimmer et al 2017). The laser pulses were delivered onto the Recently, a nanotip has been combined with a velocity map
sample most often with reflective optics, such as off-axis imaging spectrometer (Bainbridge and Bryan 2014). In order
parabolic mirrors or Schwarzschild-type microscope objec- to use nanotips for electron source applications, they can be
tives to prevent chromatic aberration and dispersion effects. integrated into electron optical imaging systems, both on a
Overlapping the nanotip apex with the laser focal spot can be macroscopic (Hoffrogge et al 2014) and microscopic scale
achieved with the help of 3D nanopositioning translation (Lüneburg et al 2013, Bormann et al 2015, Storeck
stages. et al 2017). Typically, a negative bias voltage VDC is applied
Most experiments are carried out in ultrahigh vacuum, between nanotip and detector, leading to a static electric field
helping to avoid surface contamination of the sample and of magnitude ∣FDC∣ = ∣VDC∣ (k r r ) at the tip’s apex. The
allowing the operation of electron multipliers. An exception geometry-dependent field reduction factor kr ranges, typi-
are nano-devices in ambient air where the gap between the cally, from 5 to 15 for nanotips.
sample and detection electrode is less than 5 μm, sufficient for
electron diffusion to the detector (Rybka et al 2016, Putnam
et al 2017). Nanotips from some materials can be produced 4. Photoemission mechanisms
from monocrystalline and polycrystalline wires by electro-
chemical etching, for example, tungsten tips (Klein and The near-field induced at nanotips and other nanostructures
Schwitzgebel 1997) and gold tips (Neacsu et al 2005b, Eisele opens the door to strong-field photoemission with field
et al 2011). Tungsten and to a lesser degree gold enable the strengths in excess of 1 V Å−1. This is the central theme of
use of field emission microscopy (Fursey 2005) and field ion this article, as it enables strong-field physics. In analogy to
microscopy (Müller 1965, Panitz 1982) in order to char- gas-phase atomic photoionization, photoemission from
acterize the field emission properties and the atomic-scale metallic surfaces obeys the intensity-frequency scaling of
crystallographic surface structure of the tip apex. Field eva- Keldysh’s theory (Bunkin and Fedorov 1965, Keldysh 1965).
poration (Tsong 1990) and resistive flash heating are two The dimensionless Keldysh parameter γ is given by
techniques to clean the tip in situ. State-of-the-art nanofabri-
cation techniques such as electron-beam lithography, focused
W
ion beam milling and epitaxial deposition significantly extend g= , (2 )
the range of materials and geometries. For example, recent 2Up

8
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

where the work function W for the least bound electrons in the
solid replaces the binding energy of the electron in the atom.
The ponderomotive energy Up
e2F02
Up = (3 )
4 mw 2
describes the mean quiver energy of a free electron (electron
mass m and charge e) in a monochromatic light field of
amplitude F0 and angular frequency ω. For the case of a Figure 5. Emission mechanisms assisted by a static field.
nanotip, the field strength F0 is replaced by the corresponding (a) Photoemission assisted by the Schottky effect. The work function
near-field enhanced strength at the tip apex, Feff=ξ·F0. is lowered by ΔWs due to the Schottky effect, enabling
The Keldysh parameter separates two limiting photoemission photoemission at photon energies below W. (b) Photofield emission.
regimes, the multiphoton regime (γ ? 1) and tunneling After gaining the energy of a photon, an electron is emitted by field
emission. (c) Thermally enhanced field emission. The laser pulse is
regime (γ = 1). In the multiphoton regime, the action of the heating the conduction-band electrons, resulting in field emission of
light field on the electrons at a metal surface can be treated excited electrons.
within a perturbative approach. In the tunneling regime, this
approach inevitably breaks down since the laser field strength
is starting to compete with the strength of the field binding the electron gun designs employ this mechanism to localize the
electrons within the metal. In the following, we will survey emission to the tip apex (see section 10.2). The static field can
important photoemission mechanisms occurring at metal even be used to shape the energy distribution of the emitted
nanotips and discuss their signatures in experimental data. electron pulses by tuning the work function in order to
minimize chromatic electron propagation effects (Hoffrogge
4.1. Photoemission assisted by a strong static electric field et al 2014, Ehberger et al 2015).
In contrast to extended flat metal surfaces (and to nano-
particles in the gas phase), metal nanotips offer the possibility 4.1.2. Photofield emission. Photofield emission, also called
of also applying a static electric field sufficiently strong for photo-assisted field emission, consists of two steps (Lee 1973,
field emission. At the tip apex, local field strengths of up to Radoń 1998). First, electrons are excited by single-photon
−2 GV m−1 can be easily attained and provide an additional absorption to unoccupied conduction band states. Second,
control knob not available in other geometries. The static field field emission induced by the static field leads to emission of
can also be made slowly time-dependent by applying voltage these electrons from the metal surface (see figure 5(b)). The
pulses (nanosecond time scale, Ganter et al (2008)) or THz photofield emission rate depends both linearly on laser
fields (picosecond to femtosecond time scale, see, e.g. intensity and nonlinearly on the static field strength at the
Wimmer et al (2014), Herink et al (2014), Li and Jones surface described by the Fowler–Nordheim (FN) theory
(2016), Wimmer et al (2017)). Before discussing multiphoton (equation (5)). The work function W has then to be replaced
and tunneling photoemission, we will sketch three emission by a reduced barrier height W−ÿω (Lee 1973, Hommelhoff
mechanisms that are enabled by the static field: photoemis- et al 2006c). The resulting photoelectron spectrum is a
sion assisted by the Schottky effect, photofield emission and convolution of the laser spectrum with the projected surface
thermally enhanced field emission. density of states along the laser polarization axis and with the
tunneling probability from the field emission step (Rethfeld
4.1.1. Photoemission assisted by the Schottky effect. The et al 2002). Photofield emission is particularly sensitive to
photoelectric effect is conceptionally the simplest and most electronic decoherence effects in the metal. Electron–electron
fundamental photoemission mechanism. Accordingly, the and electron–phonon scattering inherent to metals and space-
energy carried by a single photon must be larger than the charge effects can strongly influence the electron momentum
work function of the metal, requiring light in the deep UV spectrum on time scales larger than 10 fs (Rethfeld et al 2002,
spectral region for most materials (Einstein 1905). At metal Yanagisawa et al 2011, Wendelen et al 2013, Yanagisawa
nanotips, applying a static electric field leads to a significant et al 2016). Photofield emission is not restricted to single-
reduction of the potential barrier (see figure 5(a)). This effect photon absorption; recently, also two-photon absorption and
is known as the Schottky effect, here occurring at the metal- subsequent field emission was found at tungsten tips
vacuum interface (Schottky 1914). The effective decrease (Yanagisawa et al 2011, Yanagisawa 2013).
ΔWs of the work function is given by
∣e∣3 ∣FDC∣ 4.1.3. Thermally enhanced field emission. Thermally
DWs = - . (4 ) enhanced field emission is also a two-step process (Lee
4p 0
1973, Kealhofer et al 2012) (see figure 5(c)). Compared to
photofield emission, the first step is of a slightly different
The Schottky effect enables single-photon (and two- nature. Laser light strongly excites electrons within the metal,
photon) photoemission at photon energies well below the creating a non-equilibrium electron distribution. For a
work function as ΔWs can reach up to −2 eV. Laser-triggered thermalized system of conduction-band electrons, this

9
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

corresponds to heating to temperatures of the order of 1,000 K


(Lisowski et al 2004). During and after heating of the electron
gas, electrons transiently occupy states above the Fermi level.
These electrons can undergo field emission if a strong static
electric field is applied to the tip. On the time scales of several
hundred femtoseconds to picoseconds after the excitation,
electron–phonon scattering sets in, leading to a full
thermalization of the metal (Vurpillot et al 2006, Gault
et al 2007). The heating effect depends on many quantities
such as laser average power and peak intensity, pulse duration
and material properties, and competes with all other emission
mechanisms. According to the detailed investigation by
Kealhofer et al (2012), the (electronic) thermal conductivity
is of particular importance. For instance, hafnium carbide,
despite its high melting point and due to its poor electronic
thermal conductivity is much more prone to thermal effects
than metals such as gold and tungsten. Thermally enhanced
field emission has been found to be highly nonlinear in the
laser intensity, more than the usual multiphoton scaling, and
exhibits a strongly localized emission pattern resembling field
emission (Kealhofer et al 2012).

4.2. Multiphoton and above-threshold photoemission

In close analogy to multiphoton ionization in the gas phase, Figure 6. Strong-field above-threshold photoemission. (a) Count
solid surfaces enable multiphoton photoemission (MPP). Its rate as a function of energy for different incident peak intensities
main experimental signature is the power-law scaling of the (W tip, 800 nm). From bottom to top, the intensities are
emission rate w with intensity I0, w µ I0n min , where nmin is the I0 = {1.9, 2.3, 2.8, 3.2, 3.7, 4.2, 4.6} ´ 1011 W cm-2 . Color sym-
minimum number of photons required to overcome the work bols mark the features analyzed in (c). Inset: illustration of ATP from
function. MPP is associated with Keldysh parameters γ ? 1. a metal surface. (b) Dependence of the total emission current on
intensity in a double-logarithmic plot. The red line is an MPP power-
Many experimental studies at metal nanotips demonstrated law fit to the data, revealing an effective nonlinearity of n≈3.1
MPP, for instance Ropers et al (2007a, 2007c), Barwick et al (three-photon process). (c) Positions of n=4 and n=5 peaks
(2008), Tsujino et al (2008), Hilbert et al (2009) and Yana- (squares and diamonds) and two neighboring minima (circles and
gisawa et al (2011). Only recently, an experiment identified triangles) as function of intensity. The slopes of the linear fit curves
three-photon MPP by a continuous wave laser field, exploiting are in the range of −1.2 to -0.2 eV (1012 W cm-2). Adapted figure
with permission from Schenk et al (2010), Copyright (2010) by the
the strong near-field enhancement at metallic nanostar struc- American Physical Society.
tures (Sivis et al 2018). MPP is usually mediated by virtual
intermediate states or by resonant states. Resonances allow for
coherent control scenarios, where the spectral amplitude and orders appearing in the spectral measurements (figure 6(b)).
phase of a light field controls the spectral amplitude and phase ATP at nanotips and other nanostructures has been measured
of an electron wavepacket (see section 10.1). with spectral resolution in many systems and wavelength
Moving towards higher intensity and lower γ, substantial regimes, many of them in the strong-field and tunneling
contributions from higher multiphoton orders n>nmin are regimes. Prominent examples are Herink et al (2012) (Au tip,
expected. More photons than are required to free an electron 800 nm to 8 μm), Park et al (2012) (Au tip, 1.5 μm), Dombi
are absorbed and ATP takes place, first observed at flat et al (2013) (Au nano-array, 800 nm), Swanwick et al (2014)
metallic surfaces (see Luan et al 1989). Detecting ATP (n-doped Si tip array, 800 nm), Bionta et al (2014) (W tip,
requires spectrally resolved photoelectron measurements 800 nm), Bionta et al (2016) (Ag tip, 400 and 800 nm),
because the photocurrent-intensity power law will be domi- Förster et al (2016) (W tip, 1.55 μm) and Li et al (2017)
nated by the lowest multiphoton order. Figure 6(a) displays (CNT array, 820 nm). Most of these experiments, however,
the first observation of ATP at a nanotip, recorded at different do not resolve the ATP peaks; possible reasons include lim-
incident light intensities (W, λ=800 nm) (Schenk et al ited spectral resolution, space-charge broadening and thermal
2010). With increasing intensity more ATP peaks appear, contributions to the emission, surface impurity of the
corresponding to the number of absorbed photons. The employed emitters, a broad distribution of initial states in
spectrum decays exponentially with increasing kinetic energy, energy in the material or intensity averaging due to pulse
as expected from theory (see Chapter 5). The dependence of envelope effects in the strong-field regime. If photon orders
the total photocurrent on intensity shows that three-photon are not resolved, care must be taken to distinguish the prompt
MPP dominates for all intensities, overshadowing the higher and coherent ATP from other emission processes.

10
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

4.3. Onset of strong-field effects

At even higher intensities and lower Keldysh parameters, the


simple picture of MPP and ATP breaks down. Continuum
effects gain importance, in particular the dynamics of the
liberated photoelectron in the laser field, leading to shifts of
ATP peaks and channel closings. The energy of each ATP
multiphoton order reads: Ekin = nw - W - Up  0 . In
analogy to atomic systems (Bucksbaum et al 1987), the peaks
shift towards lower energies and more and more photon
orders will disappear (channel closing). A simple explanation
is provided by the AC Stark effect (or light shift). During the
presence of the laser pulse, the continuum states are field-
dressed and upshifted in energy by the ponderomotive energy
Up (Mulser et al 1993). This energy is lost after the laser pulse
has ended. Strong-field effects are visible in figure 6(a). For
increasing intensity, the peaks shift to lower energy and the
lowest peak (n=3) finally disappears. The shift of the
spectral features is captured quantitatively in figure 6(c).

4.4. Tunneling photoemission

The occurrence of more and more channel closings leads to


strong deviations from the MPP power law. The current-
intensity relation turns into a tunneling rate scaling expo-
nentially with field strength (see equation (7), Chapter 5), a
hallmark of optical tunneling emission. Indications for the
multiphoton-to-tunneling transition at a metal surface were
observed by Tóth et al (1991) and Dombi et al (2010); the
first observation at a nanostructure was reported by Bormann
et al (2010) at a gold tip with a low-repetition-rate amplified
Ti:sapphire laser system. Figure 7(a) shows the main result of
their study: a clearly resolved ‘kink’ in the scaling of the
Figure 7. Transition from the multiphoton to the optical tunneling
photocurrent with the pulse energy. A quantitative analysis regime. (a) Hallmark of the transition is the ‘kink’ in the current-
reveals a pronounced decrease of the effective nonlinearity intensity (pulse energy) scaling (here: Au tip, 830 nm). The slope in
from n∼5 to n∼1 (figure 7(b)); the authors argued that the double-logarithmic plot is changing rapidly from a multiphoton
space-charge saturation can be ruled out since the solid angle power-law scaling to a tunneling rate behavior (blue circles:
of emission measured with an MCP detector does not increase experimental data; dashed black line: multiphoton power law; red
curve: SFA calculation). Insets: single-shot images of the emission
with pulse energy. Figure 7(c) shows the result of a calcul- patterns. (b) Dependence of the effective nonlinearity (red squares)
ation using the strong-field approximation (SFA, see and solid angle of emission (green circles) on pulse energy.
section 5.3) including field penetration into the metal, clearly (c) Results of an SFA calculation including field penetration (total
reproducing the transition and elucidating its origin in the emission rate: solid black curve; selected individual channels:
progressive closing of more and more channels. colored curves). Increasing the intensity (decreasing γ) leads to the
closing of more and more channels, resulting in the ‘kink’. Adapted
The transition marked by the ‘kink’ has been observed figure with permission from Bormann et al (2010), Copyright (2010)
and studied in a wide range of systems. Among them are by the American Physical Society.
Keathley et al (2013) and Swanwick et al (2014) (n-doped Si
tip array, 800 nm), Hobbs et al (2014) (Au nano-array,
800 nm), Piglosiewicz et al (2014) (Au tip, 1.65 μm), Kusa observation of a saturation of the intensity dependence of the
et al (2015) (Au nano-array, 3 ... 10 μm), Rybka et al (2016) current by Piglosiewicz et al (2014) is noteworthy but
(Au nano-gap device, 1.5 μm), Putnam et al (2017) (Au nano- remained unexplained. A recent study presented by Keathley
array device, 1 μm), Li et al (2017) (CNT, 820 nm) and et al (2017) reveals finer structures in the scaling curve related
Keathley et al (2017) (n-doped Si tip array, 800 nm). In some to single-channel closing, as predicted by the SFA (Yalunin
studies, including the pioneering work (Bormann et al 2010), et al 2011) (see also figure 9 in section 5.3). The authors also
the transition is quite sharp, unlike the predictions of the SFA. hint at the importance of the initial density of states dis-
Incorporating the penetration of the laser field into the metal tribution for the shape of the transition. Recently, signatures
into the SFA, as suggested by (Bormann et al 2010), can of the transition have been observed in a very different sys-
explain this sharpening; additional processes limiting the tem, a bulk semiconductor nano-device (Paasch-Colberg
emission rate can be depletion or space-charge effects. The et al 2016).

11
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

5. Theoretical methods atoms and may reach the detector after a second acceleration
cycle in the light field, leading to higher asymptotic kinetic
A more detailed understanding of the processes involved in energies (step 3).
laser-nanotip interactions requires the development of theor- The tunneling step 1 is the key non-classical ingredient
etical tools which are discussed in this section. Due to the which renders the three-step model semi-classical rather than
design of typical experimental setups (Chapter 3) one is classical. Its description can be traced back to the FN theory
confronted with a true multi-scale problem: the ground-state for field emission in strong static electric fields (Fowler and
electronic wavefunction typically extends over distances a of Nordheim 1928, Nordheim 1928). In this seminal work, the
one to a few Ångstroms, the quiver amplitude α of the lib- surface potential of the solid was approximated by a potential
erated electron in the enhanced near-field may reach step function (deviations of the surface potential from the step
several nm. The tip radius is about R∼10 nm and the laser function were considered negligible) of height EF+W, with
wavelength λ is of the order of 1 μm. Associated time scales EF the Fermi energy of the electron gas and W the work
range from a v F ~ 50 as for the electrons in motion (vF K function of the material. Superimposed with the potential of
Fermi velocity) over the laser period Topt∼2.7 fs at a the electric field along the surface normal, VDC=FDCz, the
wavelength of λ=800 nm to the duration of the laser field-dependent current jFN was found to be proportional to
pulse τ5 fs.
⎡ 4 2 · W3 2⎤
Considering this range of length and time scales, mod- jFN µ FDC
2
exp ⎢ - ⎥, (5 )
eling the interaction of laser pulses with metallic nanotips ⎣ 3∣FDC∣ ⎦
appears to be a formidable task. However, these widely dis-
which can be directly converted into a field-dependent
parate scales allow for a separation of the problem into a
emission rate. A more sophisticated version of the equation
mesoscopic part for the propagation of the laser pulse in the
also accounts for the presence of the image-force poten-
presence of the metal tip (solution of Maxwell’s equations;
tial(see, e.g. Forbes 2006). Applicability of this simple esti-
see above) and a microscopic part for the simulation of the
mate to time-dependent laser fields is far from obvious and is
interaction of the local electric light field with the electronic
well justified only in the adiabatic limit. For short tunnel
system. This is because of the negligible effect of the laser-
distances and sufficiently slowly varying fields, the static
induced electron currents (∼1 electron per pulse) on the
approximation provides a good estimate for, now time-
propagation of the laser pulse. The solution of Maxwell’s
dependent, tunneling rates. Keldysh succeeded in quantifying
equations can therefore be directly used as independent input
‘short’ and ‘sufficiently slowly’ by introducing the parameter
to any microscopic model for the photoelectron emission.
named after him (see equation (2)), which can be alternatively
In the following, we will briefly review the most fre-
expressed as
quently used methods, which are capable of accounting for
recent experimental observations at least on a qualitative and (W ∣F0∣) 2W
often even on a quantitative level. The level of sophistication g= , (6 )
Topt 4p
ranges from semi-classical approaches including the classical
trajectory Monte Carlo (CTMC) method to solutions of the relating two time scales, namely the classical time it takes an
time-dependent Schrödinger equation (TDSE) on a single- electron with velocity 2W to pass through a tunnel of width
particle level to TDDFT as many-body description on a mean- W ∣F0∣ with the characteristic time of the oscillating field, i.e.
field level. Representative examples of results from different the optical period Topt. For γ=1 the barrier may be assumed
methods and comparison with experimental data will be to be static and equation (5) or a variant thereof can be used.
presented in later chapters. In this chapter, all equations are Typically, this condition is not fulfilled in laser-nanotip
given in atomic (Hartree) units (a.u.). experiments. The use of equation (5) can therefore only be
justified a posteriori when comparing with experimental data.
For example, data from recent experimental studies of pho-
5.1. (Semi-) classical trajectory simulations
toemission by two-color laser pulses strongly deviates from
In the context of laser-nanotip interactions, (semi-) classical predictions based on tunneling rates. Therefore, alternative
trajectory calculations closely mirror those employed in laser- emission processes must be involved (Förster et al 2016). For
atom interactions, the most prominent of which is the well atomic gases, a closed-form tunneling rate taking the time
known three-step model of high-harmonic generation in laser- variation of the external field into account has been proposed
atom interactions. Accordingly, electrons tunnel through the for nonadiabatic tunneling (Yudin and Ivanov 2001). An
barrier of the combined surface and laser potentials around analogous analysis for emission from metal surfaces is still
the time of the maxima of the electric field (step 1). The missing. In most simulations, equation (5) or static emission
liberated electrons are then accelerated in free space and rates derived for atoms, e.g. the Perelomov–Popov–Terent’ev
undergo a quiver motion. Depending on the phase of the (Perelomov et al 1966), Ammosov–Delone–Krainov (Ammosov
driving field at the moment of appearance at the tunneling exit et al 1986) or Delone–Krainov (Delone and Krainov 1994) rates
and the initial momentum, electrons will be either directly are used. For laser fields linearly polarized in the direction of
emitted or are driven back towards the surface after an the tip axis and for an electron in the Fermi gas with a
excursion of about the quiver radius α (step 2). A fraction of binding energy in the range of W  Wi  W + EF , application
the latter electrons is eventually (back-) scattered at surface of equation (5) for a ‘slowly’ varying laser field gives rise to a

12
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

time-dependent emission rate gives information on the momentum distribution of photo-


electrons upon release from the surface.
⎡ - C2 · W 3 2 ⎤
w (Wi , t ) = f [F (t )] Q [ - F (t )] exp ⎢ i
⎥, (7 ) For a moderate intensity of the IR laser field of
⎣ ∣F (t )∣ ⎦ I0 » 5 ´ 1011 W cm-2 at λ=800 nm well below the
damage threshold, the quiver amplitude in the enhanced near-
where f [F (t )] is a field-dependent prefactor to the dominant field, a = x · F0 w 2 » x · 1 a.u. is still small compared to the
exponential and C2 a constant. The Heaviside Θ-function characteristic distance over which the near-field enhancement
ensures that, unlike for atoms, emission takes place only during decays. Therefore, one can simulate the trajectories in a
laser half-cycles with negative field, i.e. when the barrier is spatially homogeneous enhanced field ξ·F(t).
lowered toward the vacuum side. It is important to note that in An electron emitted at the maximum of the field is
this context Ekin = W + EF - Wi is the kinetic energy of a free accelerated first towards vacuum for a quarter of an optical
electron in the direction of the surface normal. In simple esti- cycle. When the direction of the field reverses it is decelerated
mates only electrons from the Fermi edge with Wi=W are and finally driven back towards the surface which it reaches
considered. For an improved simulation, emission rates w (Wi, t ) close to the end of the next half-cycle with a kinetic energy of
are weighted with the projected density of states D (Ekin ) or, up to 3.17 Up (Milošević et al 2006).
more accurately, the projected surface density of states (SDOS). While this is the maximum energy for recombination
The simplest and most popular choice is the SDOS of a free- and, hence, high-harmonic generation, for electron emission
electron gas, D (Ekin ) µ(EF - Ekin ). the third step involves elastic scattering at the surface. For
The tunneling process not only governs the emission rate scattering angles close to Δθ≈π the momentum direction is
equation (7) but also the momentum and spatial distributions, impulsively reversed, allowing now for an additional accel-
i.e. the phase-space distribution of the liberated electron. This eration in the complete subsequent half-cycle and leading to
distribution provides the initial conditions of the subsequent final energies of up to 10.007 Up (second classical cut-off).
classical trajectory calculation. A wide variety of choices of The time evolution of the wavefunction of a free electron
initial conditions are available. in momentum space from emission time t0 to rescattering time
As for tunneling emission from atoms, position and ts is given by Volkov (1935) as
momentum distributions at the tunnel exit are a matter of
debate and cannot be reconstructed unambiguously. Estimates y ( p , ts) = exp [ - iSp (ts, t0)] y ( p , t0) , (9 )
for parallel and perpendicular momentum distributions are
available for atoms (e.g. Delone and Krainov 1994, where the Volkov phase is given by the classical action int-
Popov 1999), but are missing for metal surfaces. Often, the egral,
simple ansatz for the position at the geometric tunnel exit, ts 1
z 0 = W ∣F (t )∣ and p0=0 is used (Krüger et al 2011, Sp (ts, t0) = òt 0 2
[ p + A (t )]2 dt. (10)
2012a). More sophisticated approaches include corrections of
the position z0 due to the presence of the image potential, In a semi-classical approximation for the free-particle pro-
-1 4z , and using a Gaussian momentum distribution for the pagation, the classical trajectories are endowed with phases
emitted electron, thereby approximately accounting for the given by the classical actions. Consequently, the Volkov
position-momentum quantum uncertainty. phase Sp (ts, t0 ) (equation (10)) enters also the semi-classical
In step 2, the electron trajectory can now be easily description of strong-field dynamics, which allows to selec-
simulated numerically by solving Newton’s equation of tively incorporate quantum effects (Krüger et al 2011,
motion for a charged particle in the combined fields of the 2012b). This can be viewed as a partial sum over those paths
surface and the external laser pulse, contributing to the Feynman path integral that do not involve
dVsurf (z) classically forbidden trajectories.
p = - - F (z , t ) , (8 ) The quantitative importance of the third step is controlled
dz
by the backscattering cross-section. For the latter, a variety of
where p is the momentum along the surface normal and F(z, t) approximations are in use ranging from hard-wall scattering
is the effective near-field discussed in Chapter 2. (s  ¥ , p  -p), which is sufficient to reproduce the
Depending on the time of emission and their initial plateau and cut-off energy of the photoelectron spectrum
momentum, electrons may escape from the tip region and (Krüger et al 2011), to full quantum mechanical doubly
reach the detector directly or they may be driven back to the differential elastic scattering probabilities (cross-sections) for
surface of the solid (indicated in figure 8 by the small light surface atoms modeled by muffin-tin potentials, which also
blue wavepackets). In the former case, a Gaussian momentum allow to reproduce the relative heights of direct peak and
distribution at the tunnel exit will directly result in an expo- plateau (Salvat et al 2005, Wachter et al 2012). In the energy
nentially decreasing low-energy spectrum reaching up to range considered here, the elastic scattering probability for
about 2Up (Milošević et al 2006) with a width (1/e-intensity) electrons with Ekin20 eV off heavy atoms can reach
directly related to the width of the momentum distribution at 10%–20%.
the tunnel exit. Comparison between experimental and Based on the three steps discussed above, the main fea-
simulated energy spectra in the low-energy region therefore tures of the energy spectrum of photoelectrons emitted in

13
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Figure 8. Electron dynamics after tunneling. (a) Orange line: laser electric force, blue: direct wavepackets, light blue: rescattered
wavepackets, dashed lines: trajectories suppressed due to screening of the electric field at the surface of the tip. (b) Exponentially decreasing
spectrum due the first direct wavepacket ∣1ñ (blue filled curves). (c) The plateau forms due to rescattering of the wavepacket ∣2ñ (green filled
curves). (d) Repetition of the electron emission after one optical cycle Topt=2π/ω gives rise to interference fringes in the energy spectra
between wavepackets ∣1ñ and ∣3ñ. (e) The full spectrum also includes interferences between rescattered wavepackets ∣2ñ and ∣4ñ and closely
resembles the experimental spectrum (pink). Reproduced from Krüger et al (2012b). © IOP Publishing Ltd and Deutsche Physikalische
Gesellschaft.CC BY-NC-SA 3.0.

ultrashort laser-nanotip interactions can be reproduced: the nanotips (SMMNs), accounting for quantum effects. Con-
exponential drop at low energies and the formation of a sidering that electron emission is confined to the tip apex
plateau with its carrier-envelope phase-dependent cut-off (R∼10 nm), which is much larger than the characteristic
energy (Krüger et al 2011, Wachter et al 2012). In addition, length of the electron wavefunctions (Fermi wavelength
more complex scenarios can be simulated, in particular those λF∼0.1 nm), we find that translational symmetry of the
including the spatial decay of the nano-optical near-field in electron subsystem is approximately conserved in the surface
one or more dimensions. However, the notion of rescattering plane, allowing for a quasi-1D treatment along the surface
breaks down if the quiver length is of the order of the near- normal z.
field decay length or larger (see Herink et al (2012), Park et al For a monochromatic vector potential, A (t ) = -(Feff w )
(2012), Yalunin et al (2013), Piglosiewicz et al (2014), sin (wt ), integral (10) can be solved,
Echternkamp et al (2016b); and Chapters 7 and 8).
⎡ p2
Sp (ts, t0) = ⎢ t + pa cos (wt )
⎣ 2
ts
5.2. Simple man’s model for nanotips Up ⎤
+ Up t - ⎥ , (11)
If interference effects such as the ATP peaks and modulations 2w sin (2wt ) ⎦ t0
in the electron spectrum are also to be included within a
rescattering model, point-like classical electrons have to be and we obtain the wavefunction emitted at time t0 as
replaced by quantum mechanical or semi-classical wave-
y ( p , t , t0) = e-p 4s 2p e-iSp (t , t0 ) eWt0 .
2
packets. Here, we formulate a simple man’s model for (12)

14
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Directly emitted electrons will therefore show a momentum 5.3. Strong-field approximation
spectrum of ∣yd ( p )∣2 = e-( p 2) sp or, equivalently, an energy
2 2

The SFA, also called Keldysh–Faisal–Reiss theory (Kel-


spectrum exponentially decaying with decay constant 1 sp2 .
dysh 1965, Perelomov et al 1966, 1967, Faisal 1973,
σp can be determined directly from comparison with exper-
Reiss 1980, Gribakin and Kuchiev 1997), successfully
imental energy spectra or estimated from theory (Popov
describes photoionization of atoms and negative ions trig-
1999).
gered by strong laser fields. It corresponds to a version of a
Those parts of the wavepacket that return to the surface at
first-order distorted-wave Born approximation. The distorted
time ts are assumed to be elastically rescattered, reversing
waves used as the final state in the photoionization process
their kinetic momentum. After conclusion of the pulse this
are the Volkov wavefunctions describing free electrons driven
part of the initial wavepacket will have a momentum spec-
by a light field (see equation (9)), where the atomic potential
trum shifted by pr = 2A (ts) - A (t0 ),
is neglected after ionization took place. The SFA to the TDSE
is applicable for calculating electron spectra and ionization
∣yr ( p )∣2 = e-[ p - pr ] 2s 2p.
2
(13)
rates for all Keldysh parameters γ, provided that the ioniz-
Choosing t0 and ts to maximize pr gives the cut-off energy ation potential is much larger than the photon energy of the
pr2 2 » 10 Up . The sum of direct and rescattered momentum driving-laser field.
spectra can account for the overall shape of the energy The SFA has been first applied to strong-field photo-
spectrum quite well (Krüger et al 2012b). emission from surfaces by Bunkin and Fedorov (1965).
Periodic repetition of emission and rescattering of Neglecting the projected SDOS, the strongly nonlinear
wavepackets in subsequent laser cycles at times t j = dependence of the emission rate on the binding energy sug-
t0 + j · Topt and coherently summing up their contributions gests that the dominant emission processes can be captured by
results in the emergence of the characteristic ATP peaks in the studying the time evolution of a single active electron at the
energy spectrum. Assuming only two wavepackets launched Fermi energy, i.e. with a binding energy equal to the work
at subsequent field maxima (t0=0, t1=Topt) yields the final function of the material. The SFA rate as a function of final
momentum spectra for the direct and rescattered parts, drift momentum p is given by a sum over all channels
corresponding to the absorption of n photons,
∣yd ( p )∣2 = e-p 2s 2p
2

⎛ p2 ⎞
´ {cd,1
2
+ cd,2
2
+ 2cd,1cd,2 cos [( p2 2 + Up + W ) Topt ]} w ( p) µ å ∣In ( p )∣2 d ⎜⎝ 2
+ W + Up - nw⎟.

(16)
n
(14)
Here, we assume a reduction of the problem to one dimension
∣yr ( p )∣2 = e-( p - pr ) 2s p
2 2
and excitation by a quasi-monochromatic laser field. The
theory correctly predicts ATP peaks enforced by the energy
´ {cr,1
2
+ cr,2
2
+ 2cr,1cr,2 cos [( p2 2 + Up + W ) Topt ]},
conservation law in the delta function. The peaks are shifted
(15) to lower energies by the ponderomotive energy Up. The SFA
with ATP maxima at Ekin + Up + W = nw . The large naturally describes channel closing, i.e. due to the ponder-
separation of the two contributions in momentum space omotive action ionization channels are successively sup-
(p0=pr ) allows for a final incoherent summation to obtain pressed as the field strength is increased. Neglecting all spatial
the momentum spectrum, ∣y ( p )∣2 = ∣yd ( p )∣2 + ∣yr ( p )∣2 , dependencies of the transition matrix elements, the coeffi-
which closely resembles experimental spectra. In fact, the cients In(p) are given by
prediction (equations (14), (15)) is sufficiently close to the 2p ⎛ ⎞
1 F
observed spectrum that it can be used to directly determine In ( p ) =
2p ò0 df ⎜ik + p - 0 sin f⎟
⎝ w ⎠
the effective field strength (the free parameter in the model)
present in the experiment. To this end, the static field also ⎧⎡ pF F2 ⎤⎫
´ exp ⎨i ⎢nf + 20 cos f - 0 3 sin (2f) ⎥ ⎬ ,
present should be taken into account. It causes an effective ⎩⎣

w 8w ⎦⎭ ⎪

decrease of the work function due to the Schottky effect (see (17)
equation (4)). It also leads to a slight decrease of the cut-off with k = 2W (see Yalunin et al 2011). The argument in the
energy, Ecutoff » 10 (Feff - ∣FDC∣)2 4w 2 (Krüger et al 2012b). exponent is related to the action. The integral can be calcu-
Related concepts have been discussed for electron lated numerically. Alternatively, the total emission rate can be
emission from solid surfaces (Yalunin et al 2011, see next obtained to a good degree of approximation using a saddle-
section). In this case, however, the microscopic (electronic point approach. The generalized Keldysh rate, which covers
subsystem) and macroscopic (pulse propagation) models all γ with exponential accuracy (Bunkin and Fedorov 1965,
cannot be separated due to the large extension of the surface. Keldysh 1965, Perelomov et al 1966, Tóth et al 1991) is
Intensity averaging over the beam spot and beam-propagation
⎧ ⎡ ⎤⎫
effects erase irreversibly most of the information about details ⎪ 2 W ⎢⎛ 1 ⎞ 1 + g2 ⎪
⎥ ⎬.
of the photoexcitation process (Lemell et al 2003, Apolonski w µ exp ⎨ - ⎜ 1+ ⎟ arcsinhg -
⎩ w ⎢⎣⎝ 2g 2 ⎠ 2g ⎥⎦ ⎪

et al 2004), highlighting the importance of studies of photo- ⎭
emission from nanostructures. (18)

15
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

At large γ, in the multiphoton regime, the total photoemission


rate scales with intensity as a power law, w µ g nmin , where
nmin is the minimum number of photons required to overcome
the binding energy. In turn, the tunneling-like behavior
(see equation (7)) at higher intensities is the consequence of a
rapid sequence of channel closings (see Bormann et al (2010),
Yalunin et al (2011) and figure 7).
Application of the SFA to electron emission from the
metal-vacuum interface requires modifications and extensions
compared to the atomic case due to the broken inversion
symmetry along the surface normal. The laser field is strongly
screened inside the metal and only acts in the vacuum half-
space. Emission only takes place into the positive half-space
away from the surface. Furthermore, inclusion of the rescat-
tering process, which provides a much stronger contribution
for surfaces than for atoms because of the extended nature of
the target, requires a treatment beyond first-order distorted- Figure 9. Comparison between different theoretical approaches.
wave perturbation theory (Milošević et al 2006). So far, most Photoemission rate as a function of Keldysh parameter γ or field
investigations have resorted to atom-like semi-classical strength (800 nm, W = 5.5 eV): perturbative SFA calculations (solid
models, as described in the previous sections, and only con- orange) are compared to a TDSE solution using the Crank–
sider electron trajectories on the vacuum side of the interface. Nicholson approach (solid green) and Floquet calculations of the
four-photon (dashed blue) and five-photon (dot-dashed blue)
Accurate total photoemission rates, however, require the amplitudes. Adapted figure with permission from Yalunin et al
inclusion of rescattering at the surface, in particular in the (2011), Copyright (2011) by the American Physical Society.
weak-field regime γ>1.
In a thorough theoretical study by Yalunin et al (2011), a numerically,
time-periodic Green’s function approach has been applied to
the problem, providing an approximate solution that includes ⎡ 1 ⎤
i¶t y (z , t ) = ⎢ - ¶2z + V (z) + F (t ) z⎥ y (z , t ) , (19)
the effect of the surface. The key idea is to consider con- ⎣ 2 ⎦
tributions to the total ionization rate due to the reflection of
waves at the surface, with the help of field-dependent for sufficiently simple (model) Hamiltonians, in particular on
reflection coefficients. This can lead to an increase of the a single-particle level when many-body effects due to elec-
ionization rate compared to the plain SFA and deviations tron–electron interactions are neglected. The initial state,
from the power-law scaling in the multiphoton regime. The ψ(z, t=0), generically the ground state of a single active
results of the modified SFA compare well with Floquet electron near the Fermi edge, is propagated according to
solutions and with a numerical integration of the TDSE (see equation (19). To recover the main features of the exper-
figure 9). A different embodiment of the SFA includes the imental results, the ‘surface’ potential V(z) has to fulfill a few
influence of field penetration into the metal, influencing the requirements: the ionization potential in the ground state has
position and shape of the multiphoton-tunneling transition to be equal to W and wavepackets driven back towards the
(Bormann et al 2010). surface should be reflected with sufficiently large probability.
For a proper description of rescattering photoelectron In its most simple form, square-well potentials are used with
spectra, a second-order SFA borrowed from atomic physics one hard-wall boundary (100% reflection probability). As
should be applied (see, e.g. Milošević et al (2006) for the refinements, the image-force potential (see, e.g. Hommelhoff
atomic case). Most importantly, the theory predicts inter- et al 2006b) and the near-field’s spatial decay (Keathley
ferences of short and long trajectories, photoelectron yield et al 2017) can be included. To achieve the best agreement
beyond the classical rescattering cut-off and a modified between simulated and experimental peak positions, the depth
position of the cut-off. In analogy to the HHG cut-off and width of the potential well are adjusted to match the
(Lewenstein et al 1994, Smirnova and Ivanov 2014), the ionization potential and peak positions.
rescattering cut-off acquires a quantum correction, which At large distances from the surface the emitted wave-
yields an asymptotic value of Ecutoff » 10.007Up + 0.538 W packet is analyzed by projection onto plane waves of energy
for γ=1 (Busuladžić et al 2006). Including this correction is E=k2/2. This can either be done by expanding the wave-
important for precise determination of near-field intensities packet into box states ∣fiboxñ with eigenenergy εi, by direct
from rescattering spectra (Thomas et al 2013, Krüger Fourier transform over the detection region,
et al 2014, Rácz et al 2017, 2018). dz -ikz
y˜ (k , t ) = ò 2p
e y (z , t ) = y˜ (k ) e-iEt , (20)
5.4. Time-dependent Schrödinger equation
where ∣y˜ (k )∣2 is the probability for detection of an electron
As an alternative to semi-classical or perturbative approaches, with momentum k, or, numerically much more efficient, using
the time-dependent Schrödinger equation can be solved the so-called sampling-point method (Pohl et al 2000). To

16
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

obtain the energy spectrum the wavefunction at some detec- atomic cores can be added, e.g. the screened soft-core Cou-
tion point zD is recorded and Fourier-transformed from time to lomb potential
frequency, 1
Vatom (z) = e- ∣ z ∣ l TF , (25)
dt iwt 1 + ∣z∣
y˜ (zD , w ) = ò 2p
e y (z D , t ). (21)
with the Thomas–Fermi screening length lTF = EF 3nbulk »
Outgoing wavepackets, k > 0 can be directly mapped onto 1 a.u.»0.5 Å .
E = k 2 2 = w , yielding the photoelectron spectrum The subsequent time evolution of the system driven by
µ E ∣y˜ (zD, E )∣2 (Dinh et al 2013). The advantage of this the external time-dependent potential is described by TDDFT
method lies in the much smaller computing box with the (Runge and Gross 1984). Similar to ground-state DFT, also
‘detector’ close to the (absorbing) boundary of the system. for the time-dependent many-electron system an effective
A solution of the TDSE for a single active electron in 1D Schrödinger-like equation for a non-interacting reference
qualitatively reproduces the ionization rate scaling and the system can be constructed that, if the exact exchange-corre-
energy spectrum of photoelectrons with very small numerical lation potential were known, would yield the exact density
effort. TDSE calculations have been successfully applied to and its time derivative of the interacting system (van Leeu-
strong-field photoemission from nanotips and other nanos- wen 1999). The time-dependent orbitals ji(r, t) are solutions
tructures (Hommelhoff et al 2006b, Krüger et al 2011, of the time-dependent Kohn–Sham equation
Yalunin et al 2011, Krüger et al 2012a, Pant and Ang 2012,
Yalunin et al 2013, Ciappina et al 2014, Zhang and Lau 2016,
Keathley et al 2017). The role of the SDOS in the solid can be
i¶t ji (r , t ) = { 1
- [ - i  + A (r , t )]2 + Veff [n (r)]
2 } ji (r , t )

simulated by the incoherent addition of spectra with different (26)


initial energies of the single active electron (Yalunin In practice, suitable approximations to the potential Veff have
et al 2013). to be employed in order to allow for a numerically efficient
solution of equation (26).
5.5. TDDFT Reduction to 1D yields for the density
n occ
To reach a higher level of accuracy, the true multi-particle
nature of the electronic target system has to be taken into
n (z , t ) = å ci∣ji (z , t∣2 , (27)
i
account. This can be done by applying DFT (e.g. Parr (1994),
with the weighting coefficients ci = ∣Ei - EF∣ p from ana-
Liebsch (1997) and references therein) for the calculation of
lytical integration over the two coordinates perpendicular to
the ground-state electron density of any material. Kohn and
the tip axis and nocc the number of propagated orbitals.
Sham have devised a computationally efficient scheme to
Equation (26) reduces to the 1D Kohn–Sham equation,
generate the local ground-state density nGS(r) from a sum of
non-interacting pseudo-single-particle wavefunctions (Kohn i¶t ji (z , t ) = {-¶2z 2 + V [n (z , t )] + Vext (z , t )} ji (z , t ) ,
and Sham 1965), (28)
n (r ) = å ∣ji (r)∣2 . (22) with the time-dependent external potential
i
Vext (z , t ) = z · [FDC + Feff · fenv (t ) · cos (wt + fCE)]. (29)
The orbitals ji(r) are solutions of the single-particle Schrö-
dinger-like Kohn–Sham equation Here, fenv (t ) is the pulse envelope function, usually of
Gaussian or sin2 shape, and fCE the carrier-envelope phase.
D
{ -
2
+ Veff [n (r)] } ji (r) = ei ji (r) , (23) The photoelectron spectrum at the sampling point zD is now
calculated as the weighted sum of equation (21) for all pro-
where the effective potential pagated orbitals,
n occ ¥ 2
Veff [n (r)] = Vion [n (r)] + Ves [n (r)] + Vxc [n (r)] (24) P (E ) µ å ci ò-¥ dt eiEt ji (zD , t ) . (30)
is a functional of the local density n(r) and includes the i

interaction between the electron gas and the ionic cores, the Using TDDFT detailed comparison between measured
interaction between electrons, and contributions from and simulated spectra becomes possible. Figure 10 illustrates,
exchange-correlation, respectively. If the correct functional for the example of a tungsten tip irradiated by a 6.5 fs laser
form of Vxc [n (r)] were known, a self-consistent scheme based pulse with λ=800 nm, the potential of such TDDFT (and to
on equations (22) and (23) would converge to the exact a lesser extent CTMC) methods to successfully simulate and
ground-state density nGS(r). In practice, approximations to analyze experimental data.
Vxc [n (r)] are sought. As an additional simplification the self-
consistency condition for the potential is often abandoned.
Instead, parametrized potentials are used (e.g. Jennings 6. Temporal dynamics of photoemission
et al 1988, Chulkov et al 1999), featuring the correct
1 4z -image tail above the surface. To account for rescattering In this chapter, we will discuss the temporal dynamics of
at the first surface layer(s), a localized potential mimicking strong-field photoemission from nanotips driven by an

17
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

et al (2016). The authors report signatures of strong space-


charge broadening, distorting the spectral and temporal profile
of the resulting electron bunch with up to 1000 electrons per
pulse at a large tungsten tip. The observation of a pronounced
low-energy peak is associated with the delayed emission of
electrons returning to the tip surface and undergoing inelastic
scattering. Another pathway to uncover delays is measuring
the interferometric autocorrelation trace of the laser pulse
using the nanotip photoemission as a nonlinear element.
Current additivity at long delays indicates prompt photo-
emission and the lack of significant excitation of electrons
participating in the photoemission process, as found for
sharper tips and shorter laser pulses (Hommelhoff
et al 2006b).
A systematic investigation of delays in the multiphoton
regime is presented by Juffmann et al (2015). Photoelectrons
produced from a tungsten tip (R∼400 nm) by 10 fs, 800 nm
Figure 10. Comparison between experimental and CEP-averaged pulses were streaked in a microwave cavity, mapping their
simulated spectra. TDDFT results (solid lines) for different emission time onto energy shifts of the photoelectrons with
intensities are compared with an experimental spectrum (symbols) 2 fs accuracy. The investigation was carried out for Keldysh
for I0 = 1.7 ´ 1011 W cm-2 and a classical simulation (green dotted parameters 19.1>γ>6.6 (using the incident field strength
line), scaled to match the plateau height of the TDDFT result with
identical intensity. Best agreement between experiment and in equation (6)). Below γ∼13 the data is clearly compatible
simulation is obtained for an effective field strength with prompt emission (figure 11). The authors speculate that
Ieff = 1013 W cm-2 (field enhancement ξ≈7.5). Adapted figure the measured delays could represent Keldysh tunneling times.
with permission from Wachter et al (2012), Copyright (2012) by the
American Physical Society.
6.2. Sub-optical-cycle photoemission dynamics
ultrashort laser pulse. The basic question posed here is: is the
photoemission prompt, i.e. synchronized with the laser field, Ab initio theory provides direct insight into the attosecond
or are there delays of more than, say, an optical cycle dynamics of strong-field photoemission, without the need to
involved? Moreover, according to the simple picture of tun- rely on drastic approximations. A TDSE calculation shows
neling, we expect bursts of photoelectrons with attosecond that the temporal profile of the surface current changes sig-
duration to occur at the crests of the driving field, even in the nificantly around γ∼1. It switches from following the light
intermediate regime around γ∼1 (Yudin and Ivanov 2001, field’s vector potential, accompanied by very little net emis-
Uiberacker et al 2007). The sub-optical-cycle confinement in sion current, to following the positive crests of the laser field
time is appealing since it enables studies of attosecond elec- itself, indicating the emergence of the tunneling regime
tron dynamics at nanostructures. Here, we will discuss (Yalunin et al 2011). The sub-cycle resolving tunneling pic-
experimental evidence for delayed and prompt photoemission ture appears to remain valid even around γ∼1, as revealed
and present theoretical predictions for the sub-optical-cycle by TDSE calculations (Hommelhoff et al 2006b, Krüger
dynamics. et al 2012a). TDDFT (section 5.5) provides insight into
temporal dynamics including multi-electron effects, in good
agreement with experimental results (see figure 10). As an
6.1. Prompt versusdelayed photoemission
example, we discuss the result for a tungsten tip irradiated by
Delays in photoemission on the femtosecond and picosecond an 800 nm laser pulse with τ=6.5 fs and an effective max-
time scales are related to an excited electron population per- imum field amplitude of Feff=0.02 a.u. (Ieff = 1.4 ´
sisting even after the pulse has interacted with the sample. 1013 W cm-2 , γ∼2). In this example, the static bias field is
Thermally enhanced field emission is an obvious example, chosen to be two orders of magnitude smaller (figure 12(b)).
revealed by measuring arrival times of the electrons at the Inside the tip (z„0), the external laser field induces
detector with respect to the timing of the laser pulse. Clear density fluctuations (surface polarization) close to the surface
evidence of femtosecond delays is provided by experiments in screening the external field in the interior. The time scale of
the multiphoton regime with relatively large tungsten nano- the variation of the driving field (T≈110 a.u. ≈ 2.7 fs) is
tips (R∼100 nm) and relatively long laser pulses (∼100 fs) slow compared to the time scale of the electronic subsystem
(Yanagisawa et al 2011, Yanagisawa 2013). Supported by of the tip given by λF/vF≈5 a.u. ≈ 125 as suggesting an
simulations that include electron–electron and electron–pho- almost adiabatic adjustment of the electron gas to the external
non scattering, the authors attribute spectral shifts of the optical field. This is confirmed by the electric dipole moment
multiphoton peaks to femtosecond delays of the electron (panel (b)), which is almost perfectly in phase with the
emission with respect to the laser-pulse peak. An extreme driving field. The density change induced at the surface is less
case of multi-electron dynamics is presented by Yanagisawa than 1% of the total density in the electron gas (figure 12(a)).

18
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

wavepackets interfere, leading to ATP peaks in the energy


spectrum of emitted electrons separated by ÿω. The slope of
the maxima of δn(z, t) (figure 12(a)) indicates the velocity of
outgoing electrons. Classically, the largest slope would cor-
respond to a kinetic energy of 10 Upindicated by the dotted
arrows for rescattered electrons reached for a pulse with
suitably chosen CEP (minus-cosine-pulse). Indeed, structures
with similar slopes are reproduced by TDDFT, identifying the
rescattering mechanism as the origin of the high-energy pla-
teau in the energy spectra.

7. Acceleration and recollision of photoelectrons


Figure 11. Emission time delay in the multiphoton regime (axis on
rhs) and microwave-induced energy shift (axis on lhs) as a function In this chapter, we discuss electron dynamics after photo-
of intensity. Delays are only observed below 75 GW cm−2 (γ∼13). emission as the electrons are driven and accelerated by the
A fit curve (blue) displays the emission time delay Δt derived from strong laser field. Electrons start to quiver in the laser field,
Keldysh theory. Adapted figure with permission from Juffmann et al
(2015), Copyright (2015) by the American Physical Society. with a quiver amplitude of a = ∣e∣Feff (mw 2 ). At nanotips,
and in contrast to gas-phase photoionization, a second length
scale, the near-field decay constant lF, plays an important role.
As discussed in Chapter 2, lF is of the order of the tip radius
R. As the electron starts moving away from the surface, it may
experience the decay of the near-field. A spatial adiabaticity
parameter δ can be defined when comparing the two length
scales (Herink et al 2012),
lF l mw 2
d= = F . (31)
a ∣e∣Feff
For d  1, the electron experiences an approximately con-
stant laser field during one laser cycle. This enables the
observation of recollision phenomena, closely mirroring
recollision in the gas phase. This picture changes once we
cross into δ<1 where the dynamics are strongly modified.
Here, the electrons ‘surf’ down the near-field even before they
are able to perform significant quiver motion opening up a
true sub-optical-cycle regime of photoemission and accel-
eration (Herink et al 2012). The research performed with
nanostructures, as presented here, is closely related to surface
Figure 12. TDDFT simulation of electron emission and rescattering. (a)
Absolute value of the induced density dn (z, t ) = ∣n (z, t ) - n GS (z )∣. plasmon-driven acceleration at metal surfaces (see, e.g. Irvine
Full vertical arrows mark maxima of the electric force. Dashed lines et al (2004), Dombi et al (2010), Rácz et al (2011), Földi et al
follow trajectories of rescattered electrons. Interferences in position (2015) and the theoretical study of Rácz and Dombi (2011),
space with increasing slopes (velocities) are the signature of the equi- who predicted effects of non-quiver electron acceleration).
spaced multiphoton peaks in the electron energy spectra. (b) Full red This chapter is divided into two sections, discussing rescat-
line: laser electric force Feff(t); dashed purple line: time-dependent
dipole moment (arb.u.); dash-dotted green line: static field FDC. tering effects (δ>1) and sub-optical-cycle accelera-
Reproduced with permission from Wachter (2014). tion (δ<1).

7.1. Electron recollision and rescattering at nanotips


Turning to the vacuum side (z>0), we find the physical
The first observation of recollision and rescattering at a metal
picture developed in Chapter 5 confirmed. Electron emission
surface was reported by Krüger et al (2011) followed by more
is concentrated around the maxima of the external field systematic investigations by Wachter et al (2012) and Krüger
indicated by vertical arrows. Electrons are then accelerated et al (2012b). Figure 13(a) shows an intensity scan of ATP
towards the vacuum (subsequent half-cycle), turn around at spectra recorded at a tungsten tip. With increasing laser
the opposite field maximum and are driven back to the surface intensity, a plateau develops in the high-energy region of the
from where a fraction of the impinging wavepacket is spectrum with the count rate barely changing as a function of
rescattered (classical trajectories indicated by dotted lines). energy. The plateau is terminated by a cut-off, which depends
Rescattering will be the focus of the next chapter. As the linearly on intensity, as predicted by the 10 Up law. The
emission sequence is repeated for every laser cycle, emitted generation of the plateau is tied to the three steps of the

19
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Figure 13. Rescattering plateau in above-threshold photoemission. (a) Intensity scaling of the rescattering plateau. From bottom to top, the
incident intensity is I0 = {0.55, 0.72, 0.89, 1.1, 1.3} ´ 1011 W cm−2 (W tip, 800 nm, 6 fs). The exponential decay (direct electrons) is
followed by a plateau (rescattered electrons). The crossing points of the linear fit curves indicate the rescattering cut-off position. Inset: cut-
off energy as a function of intensity (red line: linear fit). Reproduced from Krüger et al (2012b). © IOP Publishing Ltd and Deutsche
Physikalische Gesellschaft. CC BY-NC-SA 3.0. (b) Comparison of experimental data (Krüger et al (2011), green dots) with TDSE
calculations (Yalunin et al (2013), solid green curve). For comparison, we also show results with only a single initial state (solid black curve)
and with rescattering switched off (dashed black curve). Yalunin et al (2013) © 2013 by WILEY‐VCH Verlag GmbH & Co. KGaA,
Weinheim. (c) Plateau in strong-field photoemission from Au nanorod and bow-tie arrays (800 nm). Three classes of nanorod geometries
were used, resonant (gray), red-detuned (red) and blue-detuned (blue) with respect to the laser wavelength of 805 nm. At identical incident
intensity (I0 = 2.5 ´ 1010 W cm−2), different geometries lead to variations in field enhancement, causing differences in the achieved cut-off
energy. Reprinted (adapted) with permission from Dombi et al (2013). Copyright (2013) American Chemical Society. (d) Low-intensity
(blue) and high-intensity (green) photoelectron spectra from a Si nanotip array. Reprinted (adapted) with permission from Swanwick et al
(2014). Copyright (2014) American Chemical Society. Further permissions related to the material excerpted should be directed to the ACS.

recollision picture: sub-cycle-resolved tunneling, acceleration additional static electric field, which pushes the cut-off towards
by the laser field and finally elastic scattering, the shape of the slightly lower values, in agreement with TDDFT calculations.
plateau is a result of their interplay. Although recollision does TDDFT (Wachter et al (2012), Krüger et al (2012b); see
not necessarily need to feed on the tunneling photoemission figure 10), TDSE (Krüger et al 2011, 2012a, Yalunin et al 2013),
mechanism since it is mainly a result of field-driven dynamics (semi-)classical trajectory calculations (Krüger et al 2011, 2012a,
(steps 2 and 3), the shape of the plateau indicates that a sub- Wachter et al 2012, Dombi et al 2013, Rácz et al 2017, 2018)
optical-cycle tunneling rate with a maximum around the peak and the SMMN (Krüger et al (2012b); see figure 8) are able to
of the laser field is responsible for the emission step. Tun- reproduce the experimental data even to a quantitative degree;
neling also causes a slight displacement of the photoelectron figure 13(b) shows a TDSE calculation with excellent agreement
from the surface when it appears in the continuum, leading to with the experiment (Yalunin et al 2013). The ATP peaks are
a modified cut-off law Ecutoff » 10.007 Up + 0.538 W . smeared out, mostly due to the superposition of a broad dis-
Clear signatures of rescattering at δ?1 have also been tribution of initial states. In the TDSE the backscattering can be
observed at Au nanoarrays (Dombi et al 2013, Kusa et al 2015, switched off, leading to the disappearance of the plateau.
Rácz et al 2017), see figure 13(c), n-doped Si nanotip arrays
(Swanwick et al 2014), see figure 13(d), and Ag nanotips (Bionta
7.2. Sub-optical-cycle electron acceleration in near-fields
et al 2016). The pronounced dependence of rescattering on the
sub-optical-cycle waveform of the driving field is the subject of In order to decrease δ (equation (31)), it is convenient to move
the next chapter. Krüger et al (2012b) explored the effects of an to longer wavelengths than the often used 800 nm driving

20
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

length can be extremely small (Li et al 2017). The initial


observations have been supplemented by spectral measure-
ments with angular resolution, exploring the vectorial nature
of the near-field gradient (Park et al 2012, 2013). The sub-
optical-cycle acceleration leads to an energy distribution with
a narrower cone opening angle than the recollision mech-
anism. A recent study focused on a low-energy peak at δ∼1,
reporting an intensity-dependent shift due to a nonadiabatic
ponderomotive shift (Schötz et al 2018). Sub-cycle accel-
eration is intrinsic to THz-driven electron emission from
nanotips (Herink et al 2014). Kinetic energies of up to 5 keV
have been reported (Li and Jones 2016), opening up appli-
cations as a source of high-energy electron pulses.

8. Sub-optical-cycle waveform control of electron


dynamics

Since tunneling photoemission and electron acceleration from


nanotips are driven by the sub-optical-cycle waveform, these
mechanisms can be controlled by shaping the laser field. To
this end, delicate control over the temporal electric field of the
Figure 14. Sub-optical-cycle acceleration regime. (a), (b) Photo- laser pulses is required. Two options are frequently used: the
electron trajectories driven by ultrashort pulses depend strongly on use of ultrashort few-cycle pulses with control over their CEP
whether the quiver amplitude is (a) smaller (δ ? 1, short-wavelength fCE and the use of (longer) two-color pulses with control over
irradiation) or (b) larger (δ = 1, long-wavelength irradiation) than the relative phase of the two colors. Either method exploits
the decay length of the nano-optical near-field (white region). In (b),
the exponential sensitivity of the emission probability on the
the electron escapes the near-field before substantial acceleration
back to the surface. (c) Energy spectra as a function of wavelength field amplitude in combination with symmetry breaking at the
for approximately constant intensity (Au tip, R = 12 nm). At long metal surface. In this section, we will discuss the first option,
wavelengths, the cut-off of the spectrum is wavelength-independent the influence of the carrier-envelope phase of ultrashort laser
and solely determined by the intensity—hallmarks of sub-optical- pulses on the regimes δ>1 (rescattering) and δ<1 (sub-
cycle acceleration. (d) Cut-off energy (circles, upper panel) as a optical-cycle acceleration). To the best of our knowledge,
function of wavelength for two values of total emitted charge. The
curves represent upper and lower bounds to the cut-off as calculated two-color pulses have only been applied in the multiphoton
with classical trajectory simulations including the near-field decay, regime (see section 10.1 for more details).
corresponding to local intensities of I1 = 5.4 TW cm-2 (dashed) and
I2 = 40 TW cm-2 (solid). Dash-dotted curve shows the theoretical
8.1. Waveform control of rescattering
cut-off when neglecting the near-field (I2). (e) Spatial adiabaticity
parameters δ as a function of wavelength. Adapted by permission As discussed earlier (section 5.2), the cut-off energy of the
from [Springer Nature]: [Nature] (Herink et al 2012), Nature 483
(7388), 190. plateau depends on the kinetic energy of the electron wave-
packet upon rescattering from the tip surface. Maximum
kinetic energy requires maximum force driving the electron
field (see figures 14(a), (b) for an illustration). The first back to the surface, i.e. a strong dependence on fCE is
observation of sub-optical-cycle acceleration in a nanotip expected. On top of these classical dynamics of the laser-
near-field was presented by Herink et al (2012), carried out at driven electron, the CEP is affecting the field-cycle-resolved
sharp Au tips with a wavelength-tunable laser system. tunneling probability. Figure 15 shows the results of a CEP-
Keeping the peak field strength Feff approximately constant, resolved strong-field experiment with ∼6 fs pulses at 800 nm
they used short pulses at five different center wavelengths at Keldysh parameter γ≈2 (Krüger et al 2011). An FWHM
ranging from 800 nm–8 μm and measured electron spectra intensity pulse duration of 6 fs translates into ∼2.5 optical
(see figure 14(c)). At long wavelength, as we cross into cycles, enabling pronounced CEP effects. The count rate is
δ<1, the cut-off kinetic energy deviates strongly from the found to be strongly modulated by fCE over the whole
10 Up-law and becomes independent of wavelength. Virtually spectral range, as evidenced by modulation depths ranging
no quiver motion takes place and the acceleration occurs from a few percent in the direct part up to 100% above the
within less than an optical cycle (see figures 14(d) and (e)). cut-off (figure 15(a)).
Solely the peak field strength Feff governs the acceleration Upon closer look, two distinct phase-dependent effects
and determines the shape of the spectrum. are found in the CEP-resolved data (figure 15(b)). First, the
The effect has been further investigated in a systematic variation of the cut-off position with CEP is solely related to
study (Echternkamp et al 2016b) and is also applicable to the field-driven dynamics of the liberated electron. The
different systems such as CNTs where the near-field decay waveform defines the field that the electron experiences

21
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Figure 16. Carrier-envelope phase dependence of photoelectron


spectra. (a), (b) Experimental data (Krüger et al 2011) shown as
normalized asymmetry signal (equation (32)). Increased signal (large
cut-off energy) is shown in blue, reduced emission in red. (a) Raw
Figure 15. Waveform-controlled rescattering. (a) CEP-averaged data, (b) smoothed over 1.5 eV and 0.5 π. Absolute phase has been
photoelectron spectrum (blue solid curve) consisting of the direct adjusted to match results of the TDDFT simulation (c), (d). (c) raw
part (Ekin  4 eV) and the rescattering plateau. Green points show data, (d) smoothing as for experiment. (e) CEP-dependent spectrum
the modulation depth of the count rate variation when scanning the modeled by the SMMN. Figure from Wachter (2014).
CEP. (b) Color plot showing CEP-resolved spectra on normalized
linear color scales, defined separately for the direct and rescattered
parts. Data have been recorded over a range of 2π and extended to
4π for better visibility. Yellow circles indicate the positions of the CEP-dependent spectra (Ott et al 2013). The spectral width of
cut-offs (red solid curve: sinusoidal fit). (c) Contrast of four spectral the plateau (9 eV) gives a rough estimate of the duration of
peaks in the plateau as a function of the CEP (blue diamonds; red the temporal slit (450 as) through Heisenberg’s time-energy
solid curve: sinusoidal fit curve). Adapted by permission from uncertainty relation.
[Springer Nature]: [Nature] (Krüger et al 2011) Nature 475
(7354), 78. The experimental data have been successfully modeled
with a semi-classical three-step model including quantum
interference and a TDSE calculation (Krüger et al 2011,
between photoemission and rescattering, determining the
2012a). Comparing TDDFT results to the fCE-resolved
amount of acceleration up to the time of rescattering and then experimental data (figure 16) provides a stringent test of the
towards the detector. The second effect reflects quantum accuracy of the simulation and reveals excellent agreement
interference: the contrast of the peaks in the plateau depends between theory and experiment over the full spectral range,
strongly on the CEP ranging from below 10% (fCE ~ 0 ) to including the position of the ATP peaks and even some of the
30% (fCE∼π, see figure 15(c)). In a simplified one-electron fine structure. Figure 16 displays this asymmetry A of the
picture, the high-contrast case corresponds to a scenario normalized electron spectra as a function of fCE. The asym-
where two subsequent cycles lead to the generation of elec- metry is given by the electron spectrum divided by the CEP
tron wavepackets that end up at kinetic energies in the pla- average over all spectra,
teau. Their interference leads to spectral fringes, adding a
P (E , fCE)
temporal perspective to the notion of ATP peaks. In the low- A= . (32)
contrast case, only a single optical cycle within the pulse ò dfCE P (E , fCE)
contributes a rescattered wavepacket. Therefore, no fringes The dominant feature in the data is a large-scale structure of
are expected. In essence, we find here a transition from a bent stripes that converge towards a maximum at fCE=π
temporal double- to a single-slit scenario, fully controlled by (minus-cosine-pulse) for large electron energies. Such bent
the waveform of the driver pulse. Such scenarios have also stripes have also been observed in electron spectra from
been found in atomic strong-field photoionization (Lindner atoms (Paulus et al 2003, Milošević et al 2006) and dielectric
et al 2005), cold atom optics (Szriftgiser et al 1996) and nanospheres (Zherebtsov et al 2011, 2012) and have been
neutron interferometry (Hils et al 1998). The spectral fringe exploited to determine the CEP in single-shot measurements
spacing is related to the temporal separation of the emission (Wittmann et al 2009, Johnson et al 2011, Rathje et al 2012).
of the wavepackets through a Fourier relation, ΔE=h /Δt. A simple estimate (Chapter 5) provides an intuitive
Assuming Δt to be equal to the optical-cycle duration (2.7 fs), understanding of the CEP dependence of photoemission.
we end up with the photon energy (1.55 eV). Accordingly, the Assuming electron emission at the field maximum t0, the shift
observed tilt of the fringes (see figure 15(b)) reflects a varying of the rescattered momentum distribution pr in equation (13)
temporal separation, as confirmed by a Fourier analysis of the reduces to pr = 2A (ts) with ts determined by the solution of

22
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Newton’s equations for an initial momentum at the tunnel exit


of p(t0)=0 (section 5.1). Obviously, the resulting trajectories
and rescattering times ts are sensitive to the CEP leading to
the observed CEP-dependent shift of the high-energy part of
the electron spectra. More elaborate calculations using the
SMMN (section 5.2) lead to the asymmetry structure shown
in figure 16(e), closely resembling the experimental and
TDDFT results. Both the bent-stripe structure including the
curvature and the slope of the low-energy part resulting from
the interplay of the strongly nonlinear tunneling rate and the
rescattering process are reproduced. For fCE=π (minus-
cosine-pulse), rescattering is strongest, leading to the largest
cut-off energy and largest number of electrons at high ener-
gies (blue). The tunneling probability at the maximum pre-
ceding the minus-cosine peak is, however, smaller than, e.g.
for a plus-sine-pulse, for which, in turn, a smaller rescattering
energy is found. Overall, this interplay underlies the bent-
stripe pattern in the CEP dependence.
The near-perfect agreement between model calculations
and experimental data in combination with simulations of Figure 17. Waveform control of sub-optical-cycle acceleration.
(a) CEP-resolved energy spectra (Au tip, 1.65μm). Clear modulation
pulse propagation (phase shift of the near-field at the surface of both low-energy and high-energy cut-offs are observed, marked
induced by the nanostructure) opens up the possibility to by red and black circles, respectively. (b) Position of low-energy cut-
determine the absolute phase of ultrashort laser pulses by off (red triangles) and its high-energy counterpart (black circles) as a
recording photoelectron spectra in laser-nanotip interactions function of the CEP. Solid curves represent sinusoidal fits; the two
cut-off energies oscillate almost fully out-of-phase. (c), (d) Classical
(see section 10.1). trajectory calculation results for both cut-offs. Adapted by permis-
sion from [Springer Nature]: [Nature Photonics] (Piglosiewicz
et al 2014), Nat. Phot. 8, 37.
8.2. Waveform control of sub-optical-cycle acceleration
9. Harmonic generation
Kinematic CEP effects in the tunneling regime g  1 and
sub-optical-cycle acceleration regime have been investigated So far, we have focused on the energy spectra of the electrons
by Piglosiewicz et al (2014). The authors used CEP-stable photoemitted from a metal nanotip by ultrashort laser pulses.
16 fs 1.65 μm pulses in order to drive strong-field photo- In this section, we will discuss another consequence of
emission from a gold nanotip, with an enhanced field strength strong-field laser-matter interaction, namely the emission of
of Feff » 15 V nm-1. Figure 17(a) shows CEP-dependent radiation with photon energies of n·ÿω, a process known as
spectra. Both the low-energy and the high-energy cut-offs of HHG. HHG from atoms is well understood based on the
the spectrum depend on the CEP; their positions oscillate three-step model. Although not experimentally demonstrated,
approximately out-of-phase (figure 17(b)). Classical trajectory the possibility of plasmon-enhanced gas-phase HHG has
calculations, taking into account a very small near-field decay triggered extensive theoretical work (see, e.g. Yavuz et al
constant, match the experimental results (figures 17(c), (d)). (2012), Ciappina et al (2012), Lupetti et al (2013), Pérez-
The waveform predominantly affects the dynamics of the Hernández et al (2013), Wang et al (2017), Yuan et al (2018),
accelerated electrons. for the influence of plasmonic effects on field enhancement
Another route to explore the temporal dynamics in the for nanotips see Chapter 2). Three-step models of HHG from
sub-optical-cycle acceleration regime has been implemented bulk solids in momentum space have only recently been
by Echternkamp et al (2016b). At the long-wavelength proposed (see, e.g. Golde et al 2008, Vampa et al 2014,
frontier, the authors employed ultrashort laser pulses at Higuchi et al 2014, Vampa and Brabec 2017, Chizhova et al
8.7 μm. Then, only one optical cycle contributes to the 2017). For atoms, a high-energy cut-off at Ecutoff = Ip +
emission; a single wavepacket is generated and immediately 3.17 Up was predicted from classical trajectory analysis and
accelerated in the near-field without any further interaction confirmed in experiment. For solids and surfaces of nanos-
with the surface. With the knowledge of the dynamics of the tructures, the influence of collective effects, e.g. plasmon
electron inside the near-field, the spectral shape of the emitted excitation, on HHG is difficult to assess. A single active
photoelectrons can be directly mapped on the sub-optical- electron TDSE calculation has been performed to study HHG
cycle emission rate (see figure 18). The reconstructed rate from nanotips (Ciappina et al 2014). Here, we will use
agrees well with the FN tunneling rate (equation (5)), con- TDDFT to investigate HHG from nanotips and provide esti-
firming its suitability to describe tunneling photoemission mates for the possibility to observe the emission of high-
from metallic surfaces. energy photons in experiment.

23
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Figure 18. Reconstruction of the sub-optical-cycle emission rate. (a) Time-to-energy mapping (blue) enables the reconstruction of the
instantaneous rate from the electron spectrum ((b), driven by a 8.7 μm pulse). The reconstructed rate compares well with the FN rate (gray).
Adapted by permission from [Springer Nature]: [Applied Physics B: Lasers and Optics] (Echternkamp et al 2016b), Appl. Phys. B 122
(4), 80.

To the best of our knowledge, direct HHG from nanos-


tructures including metal tips has not been observed. The
reason for this can be found by an order-of-magnitude esti-
mate for the emission rate of HHG from a nanostructure
compared to that for HHG from a gas of atoms. The number
of gas atoms in the laser focus is typically about 1010, while
the surface of a nanotip of radius 10 nm only contains about
5000 atoms. This huge disparity in the number of potential
emitters considerably reduces the emission rate and, thus, the
likelihood for the observation of HHG from isolated nanos-
tructures. We can further estimate the order of magnitude of
the linear response at the driving frequency. Modeling the
Figure 19. Simulated HHG spectra from metallic nanotips.
metallic nanotip as a nanosphere of the same radius R as the
(a) Harmonic spectra for Feff=0.005, 0.01 and 0.018 a.u. The
tip, the induced dipole moment, dielectric response of the system to a δ-kick containing all
e-1 frequencies is shown in red. (b) Time-frequency analysis of emitted
d (w ) = R F0 » R F0, e  1, (33) radiation for Feff=0.018 a.u. The highest harmonics well above the
e+2 cut-off energy of HHG in atoms are emitted in time close to the zero-
crossings of the driving-laser field (c) when emitted electrons are
at the driving wavelength in linear response serves as a
rescattered at the surface (vertical dotted lines). Figure from
measure of its scattering power. The angle-integrated emitted Wachter (2014).
radiation calculated from the dipole acceleration is then given
by (Jackson (1999))
dotted lines in panel (b)). Note that in this (nearly) free-
4p ⎛⎜ 2p ⎞⎟4 6 e-1 2

ò dW∣Escat∣2 = F02
3 ⎝l ⎠
R
e+2
. (34) electron gas calculation the influence of the plasmon excita-
tion, in particular the extension of the plateau region by the
For typical ultrashort pulses employed in laser-nanotip plasmon energy ÿωp, is clearly visible and differs from atomic
interactions (I0 = 1.1 ´ 1011 W cm-2 , focus size ∼1.5 μm, spectra. This extension can serve as an indicator for possible
τ=6.5 fs, λ=800 nm), only about half a photon is scattered metal-surface and nanostructure effects.
on average for the incident number of photons, in our Comparison of our estimate for the linear response with
example more than 108. This sets the scale for the results the intensity of higher-order harmonics in figure 19 gives an
shown in figure 19. The total spectrum (∣d̈ (w )∣2 shown for impression of the expected photon yield. Taking for example
various intensities) features a strong linear-response peak at the 7th harmonic around 11 eV photon energy and taking the
ÿω, three orders of magnitude more intense than the plateau reduction of the intensity of the plateau compared to the
region extending up to the plasmon energy. High-energy response at the driving frequency, we arrive at a yield of
components of the spectrum are observed up to around ∼10−3 harmonic photons per laser pulse. This small signal
W + 3.17Up + wp (panel (a)), with ωp the plasma frequency can only be mitigated by an increased repetition rate of the
and Up calculated from the locally enhanced field strength incident pulsed laser beam or the irradiation of large ensem-
Feff. The appearance of even harmonics is a consequence of bles of nanostructures such as free-standing copper nanowires
the broken symmetry at the surface. Time-frequency analysis grown in polymer ion-track membranes (Maurer et al 2006)
of the emission process, i.e. a windowed Fourier transform, or nano-patterned arrays of doped silicon tips (Swanwick
relates the emission times of the most intense harmonics with et al 2014), possibly further enhanced by superradiance
the recollision times of the photoemitted electrons (vertical effects. Very recently, the irradiation of blades has come into

24
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

focus (Paschen et al 2018), enhancing the number of atoms in


the laser focus and potentially boosting HHG yield by a factor
of λ/R∼100.

10. Applications of tip-based photoemission

A wide range of applications of pulsed nanoscale electron


sources has been developed in the past ten years. Here, we
turn our attention first to applications that make direct use of
the sub-optical-cycle nature of the electron dynamics occur-
ring at the source itself (Section 10.1) and then extend our
scope to experiments and applications that employ the
coherent electron pulses subsequent to their generation
(section 10.2).

10.1. Lightwave metrology and control of electron dynamics in


time and space Figure 20. Lightwave electronics and coherent control of electric
currents. (a) Sketch of a lightwave-driven nano-gap antenna device
The work presented in the previous chapters shows that the in air. The electric field F of a CEP-stable 6 fs laser pulse at 1.3 μm
photoemission and acceleration of electrons are field-driven (red and blue curves) drives a current I between two gold
processes and can be controlled by the waveform of the nanoelectrodes separated by an 8 nm gap (inset: scanning electron
micrograph). The waveform asymmetry introduced by the CEP
driving-laser field. This enables control of electric currents on determines the magnitude and direction of the current, which is
the attosecond time scale, opening up pathways towards measured by a high-precision amperemeter. (b) Average current
‘lightwave electronics’. Here, instead of a static field or a expressed in pA and electrons per pulse as a function of the CEP (red
microwave field in conventional electronics, an optical field solid curve). The gray solid curves represent eight subsequent
controls and steers electric currents at PHz frequencies measurements used for averaging. Waveforms corresponding to
three data points are indicated. Adapted by permission from
(Goulielmakis et al 2007, Krausz and Ivanov 2009, Krausz [Springer Nature]: [Nature Photonics] (Rybka et al 2016)), Nat.
and Stockman 2014) for which single-cycle optical pulses can Phot. 10. (c) Sketch of a coherent control experiment of above-
be achieved (Manzoni et al 2015, Hassan et al 2016). In order threshold photoemission from a tungsten nanotip using two colors
to make use of the control capability, a nanoscale device is (1.55 μm and a weak admixture of its second harmonic). The
necessary since the typical speed of an electron in a metal is ω−2ω delay is the experimental control parameter. (d) Total
emission current as a function of sub-optical-cycle delay. The current
given by the Fermi velocity, which is of the order of nan- is modulated by up to 97.5%, corresponding to an almost perfect
ometer per femtosecond. Initial work towards a vacuum nano- switch-on-off scenario. Reprinted figure with permission from
gap diode driven by a 5 fs, 800 nm laser pulse was performed Förster et al (2016), Copyright (2016) by the American Physical
by Higuchi et al (2015), who approached two tungsten Society. (e) Illustration of the underlying physical mechanism: two-
nanotips with different radii to a distance of about 200 nm and path quantum interference leads to a modulation of the total current.
A resonance located at an intermediate energy level (dashed line)
measured the photocurrent crossing the gap from the sharper mediates the interference (EF: Fermi energy. Weff: Schottky-lowered
tip (serving as an anode) to the blunter tip (serving as a work function).
cathode). The time resolution is of the order of hundreds of
femtoseconds corresponding to the gap crossing time. The
first sub-optical-cycle-resolved study of a laser-driven nano- nano-electrode’s geometry. Putnam et al (2017) have reported
gap diode was performed by Rybka et al (2016). CEP-stable on an experiment where strong-field photoemission from
6 fs pulses at a central wavelength of 1.3 μm were generated nanotriangle or nanorod arrays is triggered by 10 fs, 1.2 μm
and focused on a nano-gap antenna structure fabricated from pulses and detected by a large collector electrode at a distance
gold with electron-beam lithography (see figures 20(a) and of about 5 μm. The authors find a detectable CEP-dependent
(b)). The gap was only 8 nm wide, which is why the current only for the nanotriangle array since nanorods exhibit
experiment could be performed in ambient air and at room forward-backward symmetry, while at nanotriangles a single
temperature. Scanning the CEP of the 1.4-cycle pulses leads sharp feature contributes to the photoemission process. In the
to the generation of a net current through the gap and also latter case, only every other half-cycle produces photoelec-
influences its direction and magnitude. The mechanism here trons, significantly enhancing CEP effects.
is the strong dependence of tunneling photoemission on the Symmetry breaking by the waveform of the light field
waveform of the pulse. Current pulses generated from half- can also be achieved using two-color pulses. Working in the
cycles with negative sign can exceed those generated from multiphoton regime, Förster et al (2016), Paschen et al (2017)
half-cycles with positive sign and vice versa. It is the asym- and Seiffert et al (2018) used 74 fs, 1.55 μm pulses with a
metric light field that breaks the symmetry and leads to a net weak admixture of the second harmonic and measured elec-
current flow. Waveform-induced forward-backward sym- tron spectra and total current as a function of the two-color
metry breaking can be enhanced by symmetry breaking of the delay (see figures 20(c)–(e)). They found strong sinusoidal

25
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

oscillations of the total current with a periodicity of Topt/2 Superimposing strong-field photoemission of a nanotip
and a modulation depth of up to 97.5%. The experimental with a low-frequency sub-two-cycle THz field adds a slowly
results are consistent with two-path quantum interference varying time-dependent electric field to the process (Herink
where one path includes the absorption of a second harmonic et al 2014, Wimmer et al 2014). Controlling the time delay
photon and the other includes the absorption of two funda- between the THz pulse and the strong infrared driving pulse
mental photons (see figure 20(e)). A DFT calculation revealed controls the kinetic energy distribution of the resulting elec-
a possible resonance in the SDOS, allowing for this type of tron pulse, which can be used to fully suppress photoemis-
quantum path interference and a high degree of coherent sion. A systematic experimental and theoretical study
control. Another recent experiment explored the polarization revealed the strong potential of spectro-temporal electron
and delay dependence of multiphoton photoemission driven pulse shaping using a combination of static and THz fields
by a two-color field (800 + 400 nm), but did not observe sub- (Wimmer et al 2017), in analogy to THz manipulation of free-
optical-cycle quantum interference (Huang et al 2017). electron beams (Kealhofer et al 2016).
The waveform dependence of strong-field photoemission Most experiments described in this article are concerned
and electron acceleration allows for the determination of the only with electron emission from the tip apex in the forward
CEP of ultrashort laser pulses. Nanotip experiments enable direction, but the geometry and surface structure of a nanotip
CEP sensing, but for practical applications such as the sta- allow manipulation of the spatial emission pattern by means
bilization of the carrier-envelope offset frequency or single- of laser polarization and incidence angle. Such manipulations
shot CEP measurements, the electron yield needs to be were reported by Yanagisawa et al (2009, 2010, 2013) for a
increased by several orders of magnitude, for example by tungsten tip (R∼100 nm) in the photofield emission regime.
using nanostructure arrays (Schenk et al 2011). The CEP of The interplay between the locally varying work function,
the nano-optical near-field at a nanotip is shifted by a wave- corresponding to different facets on the hemispherical tip
length- and geometry-dependent phase retardation with apex, and the local strength and direction of the near-field
respect to the CEP of the incident pulse (see section 2.2). enables a high degree of directional emission control. In
Therefore, spectrally resolved photocurrent measurements addition, emission from facets with low local work function
enables spatial double-slit experiments with polarization-
enable measurements of the local CEP up to a constant phase
controlled fringe contrast (Yanagisawa et al 2017).
shift. Rescattered electrons are very sensitive to the CEP (see
Chapter 8). Strong-field photoemission from the apex of a
nanotip has been employed to measure the spatial variation of 10.2. Laser-triggered nanotips as coherent source of electron
the local CEP inside a focused broadband 4 fs laser pulse at a pulses
wavelength of 700 nm as a function of longitudinal and radial
Pulsed coherent electron beams open up a wide range of
position. The focal phase does not follow the behavior of the
applications in ultrafast electron diffraction, microscopy and
monochromatic Gouy phase, as one could naively assume,
matter-wave quantum optics. Nanotips are highly coherent
but rather shows a strong dependence on the chromatic beam
electron sources not only when operated in DC field emission
properties, characterized to first order by the wavelength- but also when triggered by ultrashort laser pulses. The virtual
dependent Rayleigh range of the beam prior to focusing source radius reff (also called effective source size) of a
(Porras 2009, Hoff et al 2017a, 2017b). The phase retardation nanotip is defined as the radius of a virtual incoherent emitter
of the near-field with respect to the incident field has been that resembles the coherence properties of the tip. The smaller
measured in a streaking experiment employing single isolated reff, the more the tip acts like a (fully coherent) point source of
attosecond XUV pulses (Förg et al 2016), implementing electrons. In DC field emission, reff can be in the order of
attosecond nanoplasmonic streaking (Stockman et al 2007, 1 nm and even less for tips with radii of the order of 10 nm
Süßmann and Kling 2011). Electrons released by XUV (Spence 2013). The transverse coherence length is defined as
photoemission are streaked by the near-field generated by an the size of coherent illumination of a sample by the electron
ultrashort infrared pulse enabling its reconstruction. beam and is inversely proportional to reff. The coherence
Rescattered electrons also encode the strength of the light properties of laser-triggered electron beams from a nanotip
field, which led to their creation and acceleration. For δ?1 were investigated for single-photon photoemission assisted by
(see Chapter 7), the rescattering cut-off encodes the intensity the Schottky effect with kinetic energies in the order of 50 eV
of the near-field. Spectral measurements have been employed (Ehberger et al 2015). Diffraction from a CNT placed close to
to determine the near-field enhancement at nanotips (Thomas the nanotip was used to determine reff ; 0.8 nm for a tip
et al 2013, Krüger et al 2014) and nanorod and bow-tie arrays radius in the order of 10 nm, which compares well with reff ;
(Rácz et al 2017, 2018). For δ∼1, the near-field decay must 0.55 nm for DC field emission. The ratio K of transverse
be taken into account, enabling a vectorial reconstruction of coherence length to beam size was found to be 0.36—the
the near-field (Park et al 2013). In addition to spectrally highest reported value for a laser-triggered electron source.
resolved studies, measuring photoemission rates enables Comparable results have been found in an ultrafast trans-
rough estimates of the strength of the near-field, both in the mission electron microscope (UTEM) with a nanotip trig-
multiphoton regime (Ropers et al 2007c, Grubisic et al 2013) gered by two-photon photoemission (Feist et al 2017); at
and the transition regime to strong fields with its ‘kink’ 200 keV electron energy, the authors find K≈0.11. Com-
around γ∼1. pared to photocathode sources used for ultrafast electron

26
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

microscopy (Reed et al 2009), nanotip emitters offer superior et al 2016). In addition, conventional electron accelerators
coherence. Fully coherent femtosecond electron sources benefit from high-brightness sources of electron bunches.
might result from laser-driven single-atom tips. These tips end Fundamental emission mechanisms and beam properties of
in a single atom and are known to be fully coherent in DC nanostructure arrays were studied in order to increase the total
field emission operation (Fink 1986, Fu et al 2001, Hom- charge per electron bunch (see Mustonen et al 2011, Nagel
melhoff et al 2009). et al 2013, Keathley et al 2013, Vilayurganapathy et al 2013,
Laser-triggered nanotip-based electron guns require Swanwick et al 2014, Gubko et al 2014, Hobbs et al 2014,
electron optics to deliver the electrons on the sample. Point- Dong et al 2015, Li et al 2017).
projection microscopy (PPM) is the simplest electron imaging Highly coherent electron pulses from nanotips also allow
technique and relies on the nanometric size of the tip with for fundamental matter-wave quantum optics with free elec-
respect to a spatially resolving electron multiplier serving as a trons in the single-electron regime (see Jones et al (2016) for
screen (Fink et al 1990). A sample is placed very close (sub- a recent review). Notable is the test of the dispersionless
μm scale) to the tip and the electron trajectories emanating nature of the magnetic Aharonov–Bohm effect, adding a
from the tip apex lead to a magnification of the sample on the time-dependent perspective to a basic quantum phenomenon
screen of up to 106 allowing for ∼1 nm-resolution holo- (Caprez et al 2007, Hilbert et al 2011, Batelaan and
graphic imaging of single molecules in PPM (Longchamp Becker 2015). The interaction of nanotip-based UTEM elec-
et al 2017). Several pioneering experiments explored imple- tron pulses with nano-optical near-fields has enabled studies
mentations of time-resolved PPM using laser-triggered of coherent manipulation and reconstruction of electron
nanotips (Quinonez et al 2013, Bainbridge et al 2016, Müller quantum states (Feist et al 2015, Echternkamp et al 2016a,
et al 2016), with one experiment reporting the observation of Priebe et al 2017). Recently, ponderomotive acceleration of
photo-excited currents in a nanowire with 50-fs resolution an electron bunch in free space has been demonstrated with
(Müller et al 2015). At close tip-sample distances, it is the help of a laser-triggered SEM (Kozák et al 2018a, 2018b).
desirable to separate the optical excitation of the tip from that Finally, nanotips open up tests of Pauli degeneracy of free
of the sample. This can be achieved using grating-coupled electrons (Lougovski and Batelaan 2011, Jones et al 2016) as
surface plasmon nanofocusing (Schröder et al 2015a, the strong confinement of electron emission in time will lead
Vogelsang et al 2015, Müller et al 2016) or a metal-coated to much higher degeneracies than the record degeneracy of
fiber tip (Casandruc et al 2015). 10−4 achieved with a DC field emitter (Kiesel et al 2002).
The integration of a laser-triggered nanotip into an
electron gun assembly usually requires a combination of a
suppressor electrode followed by acceleration and focusing
11. Conclusion
electron optics (Paarmann et al 2012, Hoffrogge et al 2014,
Bormann et al 2015, Storeck et al 2017). Due to matter-wave
Strong-field-driven photoemission from nanotips has been
dispersion of free electrons on the way to the sample, electron
developed into a powerful small-scale strong-field laboratory,
pulse durations are limited to about 40 fs. Temporal electron
revealing for the first time fundamental attosecond phenom-
bunch compression to femtosecond durations, even up to the
ena at the surface of nanoscale solids. The versatility of
formation of attosecond pulse trains, can be achieved through
nanotips has led to a wide variety of applications, with
interactions with microwave cavities, with near-fields at
ramifications in lightwave electronics and ultrafast material
nanostructures and material discontinuities in the THz and
science. Now the way is open to apply advanced methods of
optical domain (Kealhofer et al 2016, Hassan et al 2017,
attosecond spectroscopy to the nanoscale in order to study
Kozák et al 2017, Priebe et al 2017, Morimoto and
ultrafast phenomena at surfaces and interfaces in real time.
Baum 2018) and with the field of high-intensity optical tra-
We anticipate new and exciting research in the coming dec-
veling waves (Kozák et al 2018a, 2018b). Sub-keV electron
ade, enabled by the extreme localization of light and matter
pulses from nanotips have been applied for low-energy
on the (sub-) nanometer length and attosecond time scales.
electron diffraction (Gulde et al 2014, Müller et al 2015,
Storeck et al 2017, Vogelgesang et al 2018). The integration
of a laser-triggered nano-emitter has been demonstrated in a
scanning electron microscope (SEM) (Yang 2010, Kozák 12. List of symbols
et al 2018a) and in a UTEM reaching electron energies in
excess of 100 keV (Caruso et al 2017, Feist et al 2017, • A(t): vector potential
Plemmons and Flannigan 2017, Houdellier et al 2018). • α: quiver radius
Nanotip photoemission is also explored for source • β: electron beam full opening angle
applications in electron accelerators because of its high- • c, C: constants
brightness B. The brightness is defined as B = j (pbreff )2 , • d: dipole moment
where β is the beam’s full opening angle and j is the total • D: density of states
emission current. A laser-triggered nanotip can serve as the • δ: dimensionless spatial adiabacity parameter
electron source in a multi-stage dielectric laser accelerator, • δn: induced charge density change
requiring high-brightness electron beams (Breuer and Hom- • Δf: near-field phase shift
melhoff 2013, Peralta et al 2013, England et al 2014, McNeur • Δθ: scattering angle

27
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

• ΔWs: lowering of the surface-vacuum barrier by the • vF: Fermi velocity


Schottky effect • V: potential energy
• e: electron charge • Veff: effective potential
• Ecutoff: cut-off energy • Vext: external potential
• EF: Fermi energy • Vsurf: surface potential
• Ekin: kinetic energy • w: emission rate
• Ephot: photon energy • W: work function
• ò: dielectric function • Wi: electron binding energy
• ò0: dielectric constant • ξ: field enhancement factor
• f: prefactor to tunneling rates • z: reaction coordinate along surface normal
• fenv (t ): light field envelope function • zD: detector position
• F(t): time-dependent optical near-field
• F0: electric field amplitude
• FDC: DC field
• Feff: effective field amplitude including field enhancement 13. List of acronyms
• γ: Keldysh parameter
• 2PPE: two-photon photoemission
• ÿ: (reduced) Planck constant.
• APT: atom probe tomography
• I0: cycle-averaged peak intensity
• ATI: above-threshold ionization
• Ieff: effective cycle-averaged peak intensity including
• ATP: above-threshold photoemission
field enhancement
• BEM: boundary element method
• In: coefficients for the strong-field approximation
• CEP: carrier-envelope phase
• Ip: atomic ionization potential
• CNT: carbon nanotube
• j: electric current
• CTMC: classical trajectory Monte Carlo
• jFN: field emission current
• DFT: density functional theory
• k: wave number (electron or photon)
• SDOS: surface density of states
• kr: field reduction factor
• FDTD: finite-difference time-domain
• K: ratio of the transverse coherence length and electron-
• FN: Fowler and Nordheim
beam size
• FWHM: full width at half maximum
• lF: near-field decay constant
• HHG: high-harmonic generation
• λ: wavelength
• IR: infrared
• λF: Fermi wavelength
• LSP: localized surface plasmon
• λTF: Thomas–Fermi screening length
• MCP: micro-channel plate
• m: electron mass
• MPP: multiphoton photoemission
• nmin: minimum required number of photons to overcome
• OAP: off-axis parabolic mirror
the work function
• PPM: point-projection microscopy
• n(z, t): charge density
• SEM: scanning electron microscope
• ω: angular frequency
• SFA: strong-field approximation
• ωp: plasma frequency • SMMN: simple man’s model for nanotips
• Ω: solid emission angle
• SNOM: scanning near-field microscopy
• p: momentum
• TDDFT: time-dependent density functional theory
• P: electron spectrum
• TDSE: time-dependent Schrödinger equation
• fCE: carrier-envelope phase
• TERS: tip-enhanced Raman scattering
• fi: non-interacting pseudo-single-particle wavefunctions
• TOF: time of flight
• ψ: wavefunction
• UTEM: ultrafast transmission electron microscope
• ψd: direct wavepacket
• UV: ultraviolet
• ψr: rescattered wavepacket • VMI: velocity map imaging
• reff: effective source radius
• XUV: extreme ultraviolet
• R: tip radius of curvature
• S: action
• σp: width of momentum distribution
• t: time Acknowledgments
• t0: ionization time
• ts: recollision time MKand PHacknowledge funding through ERC Con-
• Topt optical-cycle duration solidator Grant ‘Near Field Atto’, DFG SFB 953, DFG SPP
• θ: tip half-opening angle 1840 ‘QUTIF’. CL, GWand JBthank the Austrian Science
• Θ: Heaviside function Fund (FWF), project numbers SFB-041 ViCoM, SFB-049
• Up: ponderomotive energy NextLite, and doctoral college W1243, and also the COST
• v: velocity Action CM1204 (XLIC) and the IMPRS-APS for support.

28
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Calculations were performed using the Vienna Scientific Caprez A, Barwick B and Batelaan H 2007 Phys. Rev. Lett. 99
Cluster (VSC). MKacknowledges financial support from the 210401
Minerva Foundation and the Koshland Foundation. Caruso G M, Houdellier F, Abeilhou P and Arbouet A 2017 Appl.
Phys. Lett. 111 023101
Casandruc A, Kassier G, Zia H, Bücker R and Miller R J D 2015
J. Vac Sci. Tech. B 33 03C101
ORCID iDs Chaturvedi P, Hsu K H, Kumar A, Fung K H, Mabon J C and
Fang N X 2009 ACS Nano 3 2965
Michael Krüger https://orcid.org/0000-0002-6188-9179 Chizhova L A, Libisch F and Burgdörfer J 2017 Phys. Rev. B 95
085436
Christoph Lemell https://orcid.org/0000-0003-2560-4495
Chulkov E, Silkin V and Echenique P 1999 Surf. Sci. 437 330
Peter Hommelhoff https://orcid.org/0000-0003- Ciappina M F, Biegert J, Quidant R and Lewenstein M 2012 Phys.
4757-5410 Rev. A 85 033828
Ciappina M F et al 2017 Rep. Prog. Phys. 80 054401
Ciappina M F, Pérez-Hernández J A, Shaaran T, Lewenstein M,
References Krüger M and Hommelhoff P 2014 Phys. Rev. A 89 013409
Ciracì C, Hill R T, Mock J J, Urzhumov Y,
Fernández-Domínguez A I, Maier S A, Pendry J B,
Aeschlimann M, Schmuttenmaer C A, Elsayed-Ali H E, Chilkoti A and Smith D R 2012 Science 337 1072
Miller R J D, Cao J, Gao Y and Mantell D A 1995 J. Chem. Corkum P B 1993 Phys. Rev. Lett. 71 1994
Phys. 102 8606 Damascelli A, Gabetta G, Lumachi A, Fini L and Parmigiani F 1996
Agostini P, Fabre F, Mainfray G, Petite G and Rahman N K 1979 Phys. Rev. B 54 6031
Phys. Rev. Lett. 42 1127 Delone N B and Krainov V P 1994 Multiphoton Processes in Atoms
Ahn B et al 2017 APL Photonics 2 036104 (Berlin: Springer)
Ahn B, Schötz J, Okell W A, Süßmann F, Förg B, Kim S, Diddams S A, Jones D J, Ye J, Cundiff S T, Hall J L, Ranka J K,
Kling M and Kim D 2016 Opt. Express 24 92 Windeler R S, Holzwarth R, Udem T and Hänsch T W 2000
Ammosov M, Delone N and Krainov V 1986 Sov. Phys. JETP Phys. Rev. Lett. 84 5102
64 1191 Dinh P M, Romaniello P, Reinhard P-G and Suraud E 2013 Phys.
Apolonski A et al 2004 Phys. Rev. Lett. 92 073902 Rev. A 87 032514
Bainbridge A R and Bryan W A 2014 New J. Phys. 16 103031 Dombi P, Hörl A, Rácz P, Marton I, Trügler A, Krenn J R and
Bainbridge A R, Myers C W B and Bryan W A 2016 Struct. Dyn. 3 Hohenester U 2013 Nano Lett. 13 674
023612 Dombi P et al 2010 Opt.Express 18 24206
Baltuška A et al 2003 Nature 421 611 Dong C D, Swanwick M E, Keathley P D, Kärtner F X and
Banfi F, Giannetti C, Ferrini G, Galimberti G, Pagliara S, Velásquez-García L F 2015 Nanotechnology 26 265202
Fausti D and Parmigiani F 2005 Phys. Rev. Lett. 94 037601 Echternkamp K E, Feist A, Schäfer S and Ropers C 2016a Nat. Phys.
Barwick B, Flannigan D J and Zewail A H 2009 Nature 462 902 12 1000
Barwick B, Park H S, Kwon O-H, Baskin J S and Zewail A H 2008 Echternkamp K E, Herink G, Yalunin S V, Rademann K,
Science 322 1227 Schäfer S and Ropers C 2016b Appl. Phys. B 122 80
Batelaan H and Becker M 2015 Europhys. Lett. 112 40006 Ehberger D, Hammer J, Eisele M, Krüger M, Noe J, Högele A and
Bauer M, Marienfeld A and Aeschlimann M 2015 Prog. Surf. Sci. Hommelhoff P 2015 Phys. Rev. Lett. 114 227601
90 319 Einstein A 1905 Ann. Phys. 322 132
Becker W, Lohr A and Kleber M 1995 Quantum Semiclass. Opt. Eisele M, Krüger M, Schenk M, Ziegler A and Hommelhoff P 2011
7 423 Rev. Sci. Instrum. 82 026101
Berweger S, Atkin J M, Olmon R L and Raschke M B 2012 J. Phys. England R J et al 2014 Rev. Mod. Phys. 86 1337
Chem. Lett. 3 945 Esteban R, Borisov A G, Nordlander P and Aizpurua J 2012 Nat.
Bionta M, Chalopin B, Champeaux J, Faure S, Masseboeuf A, Commun. 3 825
Moretto-Capelle P and Chatel B 2014 J. Mod. Opt. 61 833 Esteban R, Zugarramurdi A, Zhang P, Nordlander P,
Bionta M R, Weber S J, Blum I, Mauchain J, Chatel B and Garcia-Vidal F J, Borisov A G and Aizpurua J 2015 Faraday
Chalopin B 2016 New J. Phys. 18 103010 Discuss. 178 151
Bisio F, Nývlt M, Franta J, Petek H and Kirschner J 2006 Phys. Rev. Faisal F H M 1973 J. Phys. B: At. Mol. Phys. 6 L89
Lett. 96 087601 Faisal F H M, Kamiński J Z and Saczuk E 2005 Phys. Rev. A 72
Bohren C F and Huffman D R 1998 Absorption and Scattering of 023412
Light by Small Particles (Berlin: Wiley) Fann W S, Storz R and Bokor J 1991 Phys. Rev. B 44 10980
Bormann R, Gulde M, Weismann A, Yalunin S V and Ropers C Farkas G and Chin S L 1985 Appl. Phys. B 37 141
2010 Phys. Rev. Lett. 105 147601 Farkas G, Horváth Z G, Kertész I and Kiss G 1971 Lettere al Nuovo
Bormann R, Strauch S, Schäfer S and Ropers C 2015 J. Appl. Phys. Cimento (1971-1985) 1 314
118 173105 Farkas G, Horváth Z G, Kertész I and Kiss G 1972 Phys. Lett. A
Bouhelier A, Beversluis M, Hartschuh A and Novotny L 2003 Phys. 39 231
Rev. Lett. 90 013903 Farkas G, Tóth C and Kőházi-Kis A 1993 Opt. Eng. 32 2476
Boussoukaya M, Bergeret H, Chehab R, Leblond B and Le Duff J Feist A et al 2017 Ultramicroscopy 176 63
1989 Nucl. Instrum. Methods Phys. Res. A 279 405 Feist A, Echternkamp K E, Schauss J, Yalunin S V, Schäfer S and
Breuer J and Hommelhoff P 2013 Phys. Rev. Lett. 111 134803 Ropers C 2015 Nature 521 200
Bucksbaum P H, Freeman R R, Bashkansky M and McIlrath T J Ferray M, L’Huillier A, Li X F, Lompre L A, Mainfray G and
1987 J. Opt. Soc. Am. B 4 760 Manus C 1988 J. Phys. B: At. Mol. Opt. Phys. 21 L31
Bunkin F V and Fedorov M V 1965 Sov. Phys. JETP 21 896 Ferrini G, Banfi F, Giannetti C and Parmigiani F 2009 Nucl. Instrum.
Busuladžić M, Gazibegović-Busuladžić A and Milošević D B 2006 Methods Phys. Res. A 601 123
Laser Phys. 16 289 Fink H-W 1986 IBM J. Res. Dev. 30 460
Calegari F, Sansone G, Stagira S, Vozzi C and Nisoli M 2016 Fink H-W, Stocker W and Schmid H 1990 Phys. Rev. Lett. 65
J. Phys. B: At. Mol. Opt. Phys. 49 062001 1204

29
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Földi P, Márton I, Német N, Ayadi V and Dombi P 2015 Appl. Phys. Hommelhoff P, Kealhofer C, Aghajani-Talesh A, Sortais Y R,
Lett. 106 013111 Foreman S M and Kasevich M A 2009 Ultramicroscopy
Forbes R G 2006 Appl. Phys. Lett. 89 113122 109 423
Förg B et al 2016 Nat. Commun. 7 11717 Hommelhoff P, Kealhofer C and Kasevich M A 2006a Proc. 2006
Förster M, Paschen T, Krüger M, Lemell C, Wachter G, Libisch F, IEEE Int. Freq. Control Symp. Expos. 1 + 2, 470–4
Madlener T, Burgdörfer J and Hommelhoff P 2016 Phys. Rev. Hommelhoff P, Kealhofer C and Kasevich M A 2006b Phys. Rev.
Lett. 117 217601 Lett. 97 247402
Fowler R H and Nordheim L 1928 Proc. R. Soc. London A 119 173 Hommelhoff P, Kealhofer C and Kasevich M A 2007 Ultrafast
Fu T-Y, Cheng L-C, Nien C-H and Tsong T T 2001 Phys. Rev. B 64 Phenomena XV ed P Corkum et al (Berlin: Springer) p 746
113401 Hommelhoff P and Kling M F 2015 Attosecond Nanophysics: From
Fursey G N 2005 Field Emission in Vacuum Microelectronics (New Basic Science to Applications (Weinheim: Wiley)
York: Kluwer Academic/Plenum) Hommelhoff P, Sortais Y, Aghajani-Talesh A and Kasevich M A
Gallagher T F 1988 Phys. Rev. Lett. 61 2304 2006c Phys. Rev. Lett. 96 077401
Ganter R, Bakker R, Gough C, Leemann S C, Paraliev M, Houdellier F, Caruso G, Weber S, Kociak M and Arbouet A 2018
Pedrozzi M, Le Pimpec F, Schlott V, Rivkin L and Wrulich A Ultramicroscopy 186 128
2008 Phys. Rev. Lett. 100 064801 Huang W C-W, Becker M, Beck J and Batelaan H 2017 New J.
Gault B, Moody M P, Cairney J M and Ringer S P 2012 Atom Probe Phys. 19 023011
Microscopy (New York: Springer) Huth F, Chuvilin A, Schnell M, Amenabar I, Krutokhvostov R,
Gault B, Vella A, Vurpillot F, Menand A, Blavette D and Lopatin S and Hillenbrand R 2013 Nano Lett. 13 1065
Deconihout B 2007 Ultramicroscopy 107 713 Inouye Y and Kawata S 1994 Opt. Lett. 19 159
Girardeau-Montaut J P and Girardeau-Montaut C 1995 Phys. Rev. B Irvine S E, Dechant A and Elezzabi A Y 2004 Phys. Rev. Lett. 93
51 13560 184801
Golde D, Meier T and Koch S W 2008 Phys. Rev. B 77 075330 Jackson J 1999 Classical Electrodynamics (New York: Wiley)
Goulielmakis E, Yakovlev V S, Cavalieri A L, Uiberacker M, Jennings P J, Jones R O and Weinert M 1988 Phys. Rev. B 37 6113
Pervak V, Apolonski A, Kienberger R, Kleineberg U and Johnson N G et al 2011 Phys. Rev. A 83 013412
Krausz F 2007 Science 317 769 Jones D J, Diddams S A, Ranka J K, Stentz A, Windeler R S,
Gribakin G F and Kuchiev M Y 1997 Phys. Rev. A 55 3760 Hall J L and Cundiff S T 2000 Science 288 635
Grubisic A, Schweikhard V, Baker T A and Nesbitt D J 2013 ACS Jones E, Becker M, Luiten J and Batelaan H 2016 Laser Photonics
Nano 7 87 Rev. 10 214
Gubko M A, Husinsky W, Ionin A A, Kudryashov S I, Juffmann T, Klopfer B B, Skulason G E, Kealhofer C, Xiao F,
Makarov S V, Nathala C R, Rudenko A A, Seleznev L V, Foreman S M and Kasevich M A 2015 Phys. Rev. Lett. 115
Sinitsyn D V and Treshin I V 2014 Laser Phys. Lett. 11 264803
065301 Kealhofer C, Foreman S M, Gerlich S and Kasevich M A 2012 Phys.
Gulde M, Schweda S, Storeck G, Maiti M, Yu H K, Wodtke A M, Rev. B 86 035405
Schäfer S and Ropers C 2014 Science 345 200 Kealhofer C, Schneider W, Ehberger D, Ryabov A, Krausz F and
Hänsch T W 2006 Rev. Mod. Phys. 78 1297 Baum P 2016 Science 352 429
Hartschuh A 2008 Angewandte Chemie International Edition Keathley P D, Putnam W P, Laurent G, Velásquez-García L F and
47 8178 Kärtner F X 2017 arXiv:1708.05265
Hassan M T, Baskin J S, Liao B and Zewail A H 2017 Nat. Keathley P D, Sell A, Putnam W P, Guerrera S, Velasquez-Garcia L and
Photonics 11 425 Kärtner F X 2013 Ann. Phys. (Berlin) 525 144
Hassan M T et al 2016 Nature 530 66 Keldysh L V 1965 Sov. Phys. JETP 20 1307
Herink G, Solli D R, Gulde M and Ropers C 2012 Nature 483 190 Kienberger R et al 2004 Nature 427 817
Herink G, Wimmer L and Ropers C 2014 New J. Phys. 16 123005 Kiesel H, Renz A and Hasselbach F 2002 Nature 418 392
Hernandez Garcia C and Brau C A 2002 Nucl. Instrum. Methods Klein M and Schwitzgebel G 1997 Rev. Sci. Instrum. 68 3099
Phys. Res. A 483 273 Kohn W and Sham L J 1965 Phys. Rev. 140 A1133
Higuchi T, Maisenbacher L, Liehl A, Dombi P and Hommelhoff P Kozák M, Eckstein T, Schönenberger N and Hommelhoff P 2018a
2015 Appl. Phys. Lett. 106 051109 Nat. Phys. 14 121
Higuchi T, Stockman M I and Hommelhoff P 2014 Phys. Rev. Lett. Kozák M, McNeur J, Leedle K J, Deng H, Schönenberger N,
113 213901 Ruehl A, Hartl I, Harris J S, Byer R L and Hommelhoff P 2017
Hilbert S A, Barwick B, Fabrikant M, Uiterwaal C J G J and Nat. Commun. 8 14342
Batelaan H 2007 Appl. Phys. Lett. 91 173506 Kozák M, Schönenberger N and Hommelhoff P 2018b Phys. Rev.
Hilbert S A, Caprez A and Batelaan H 2011 New J. Phys. 13 093025 Lett. 120 103203
Hilbert S A, Neukirch A, Uiterwaal C J G J and Batelaan H 2009 Krausz F and Ivanov M 2009 Rev. Mod. Phys. 81 163
J. Phys. B: At. Mol. Opt. Phys. 42 141001 Krausz F and Stockman M I 2014 Nat. Photonics 8 205
Hils T, Felber J, Gähler R, Gläser W, Golub R, Habicht K and Krüger M, Schenk M, Förster M and Hommelhoff P 2012a J. Phys.
Wille P 1998 Phys. Rev. A 58 4784 B: At. Mol. Opt. Phys. 45 074006
Hobbs R G, Putnam W P, Fallahi A, Yang Y, Kärtner F X and Krüger M, Schenk M and Hommelhoff P 2011 Nature 475 78
Berggren K K 2017 Nano Lett. 17 6069 Krüger M, Schenk M, Hommelhoff P, Wachter G, Lemell C and
Hobbs R G, Yang Y, Fallahi A, Keathley P D, Leo E D, Kärtner F X, Burgdörfer J 2012b New J. Phys. 14 085019
Graves W S and Berggren K K 2014 ACS Nano 8 11474 Krüger M, Thomas S, Förster M and Hommelhoff P 2014 J. Phys. B:
Hoff D, Krüger M, Maisenbacher L, Paulus G G, Hommelhoff P and At. Mol. Opt. Phys. 47 124022
Sayler A M 2017a J. Opt. 19 124007 Kruit P, Kimman J, Muller H G and van der Wiel M J 1983 Phys.
Hoff D, Krüger M, Maisenbacher L, Sayler A M, Paulus G G and Rev. A 28 248
Hommelhoff P 2017b Nat. Phys. 13 947 Kruit P, Kimman J and der Wiel M J V 1981 J. Phys. B: At. Mol.
Hoffrogge J, Stein J P, Krüger M, Förster M, Hammer J, Ehberger D, Phys. 14 L597
Baum P and Hommelhoff P 2014 J. Appl. Phys. 115 094506 Kuchiev M Y 1987 JETP Lett. 45 404
Homann C, Bradler M, Förster M, Hommelhoff P and Riedle E 2012 Kusa F, Echternkamp K E, Herink G, Ropers C and Ashihara S 2015
Opt. Lett. 37 1673 AIP Adv. 5 077138

30
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Lee M J G 1973 Phys. Rev. Lett. 30 1193 Nagel P M, Robinson J S, Harteneck B D, Pfeifer T, Abel M J,
Lee M J G, Reifenberger R, Robins E S and Lindenmayr H G 1980 Prell J S, Neumark D M, Kaindl R A and Leone S R 2013
J. Appl. Phys. 51 4996 Chem. Phys. 414 106
van Leeuwen R 1999 Phys. Rev. Lett. 82 3863 Neacsu C C, Reider G A and Raschke M B 2005a Phys. Rev. B 71
Lehr M, Foerster B, Schmitt M, Krüger K, Sönnichsen C, 201402
Schönhense G and Elmers H-J 2017 Nano Lett. 17 6606 Neacsu C C, Steudle G A and Raschke M B 2005b Appl. Phys. B
Lemell C, Tong X-M, Krausz F and Burgdörfer J 2003 Phys. Rev. 80 295
Lett. 90 076403 Nelayah J, Kociak M, Stéphan O, de Abajo F J G, Tencé M,
Lewenstein M, Balcou P, Ivanov M Y, L’Huiller A and Corkum P B Henrard L, Taverna D, Pastoriza-Santos I,
1994 Phys. Rev. A 49 2117 Liz-Marzán L M and Colliex C 2007 Nat. Phys. 3 348
Li C et al 2017 Appl. Phys. Lett. 111 133101 Nordheim L 1928 Proc. R. Soc. London A 121 626
Li H et al 2015 Phys. Rev. Lett. 114 123004 Novotny L and Hecht B 2012 Principles of Nano-Optics
Li S and Jones R R 2016 Nat. Commun. 7 13405 (Cambridge: Cambridge University Press)
Liebsch A 1997 Electronic Excitations at Metal Surfaces (New Ott C et al 2013 New J. Phys. 15 073031
York: Plenum) Paarmann A et al 2012 J. Appl. Phys. 112 113109
van Linden, van den Heuvell H B and Muller H G 1988 Studies in Paasch-Colberg T et al 2016 Optica 3 1358
Modern Optics No 8, Multiphoton Processes (Cambridge: Panitz J A 1982 J. Phys. E: Sci. Instr. 15 1281
Cambridge University Press) p 25 Pant M and Ang L K 2012 Phys. Rev. B 86 045423
Lindner F, Schätzel M G, Walther H, Baltuška A, Goulielmakis E, Park D J, Piglosiewicz B, Schmidt S, Kollmann H, Mascheck M,
Krausz F, Milošević D B, Bauer D, Becker W and Paulus G G Groß P and Lienau C 2013 Ann. Phys. (Berlin) 525 135
2005 Phys. Rev. Lett. 95 040401 Park D J, Piglosiewicz B, Schmidt S, Kollmann H, Mascheck M and
Lisowski M, Loukakos P, Bovensiepen U, Stähler J, Gahl C and Lienau C 2012 Phys. Rev. Lett. 109 244803
Wolf M 2004 Appl. Phys. A 78 165 Parr R G 1994 Density-functional Theory of Atoms and Molecules
Longchamp J-N, Rauschenbach S, Abb S, Escher C, (Oxford: Oxford University Press)
Latychevskaia T, Kern K and Fink H-W 2017 Proc. Natl Acad. Paschen T, Förster M, Krüger M, Lemell C, Wachter G, Libisch F,
Sci. USA 114 1474 Madlener T, Burgdörfer J and Hommelhoff P 2017 J. Mod.
Lougovski P and Batelaan H 2011 Phys. Rev. A 84 023417 Opt. 64 1054
Luan S, Hippler R, Schwier H and Lutz H O 1989 Europhys. Lett. Paschen T, Russell R, Naranjo B, Rosenzweig J and Hommelhoff P
9 489 2018 private communication
Lüneburg S, Müller M, Paarmann A and Ernstorfer R 2013 Appl. Passig J et al 2017 Nat. Commun. 8 1181
Phys. Lett. 103 213506 Paulus G G, Lindner F, Walther H, Baltus̆ka A, Goulielmakis E,
Lupetti M, Kling M F and Scrinzi A 2013 Phys. Rev. Lett. 110 Lezius M and Krausz F 2003 Phys. Rev. Lett. 91 253004
223903 Peralta E et al 2013 Nature 503 91
Maier S 2007 Plasmonics: Fundamentals And Applications (Berlin: Perelomov A M, Popov V S and Terent’ev M V 1966 Sov. Phys.
Springer) JETP 23 924
Manzoni C, Mücke O D, Cirmi G, Fang S, Moses J, Huang S-W, Perelomov A M, Popov V S and Terent’ev M V 1967 Sov. Phys.
Hong K-H, Cerullo G and Kärtner F X 2015 Laser Photonics JETP 24 207
Rev. 9 129 Pérez-Hernández J A, Ciappina M F, Lewenstein M, Roso L and
Marinica D, Kazansky A, Nordlander P, Aizpurua J and Zaïr A 2013 Phys. Rev. Lett. 110 053001
Borisov A G 2012 Nano Lett. 12 1333 Petek H and Ogawa S 1997 Prog. Surf. Sci. 56 239
Marinica D C, Zapata M, Nordlander P, Kazansky A K, Piglosiewicz B, Schmidt S, Park D J, Vogelsang J, Gross P,
Echenique P M, Aizpurua J and Borisov A G 2015 Sci. Adv. 1 Manzoni C, Farinello P, Cerullo G and Lienau C 2014 Nat.
e1501095 Photonics 8 37
Maurer F, Dangwal A, Lysenkov D, Müller G, Toimil-Molares M E, Plemmons D A and Flannigan D J 2017 Chem. Phys. Lett. 683 186
Trautmann C, Brötz J and Fuess H 2006 Nucl. Instrum. Pohl A, Reinhard P G and Suraud E 2000 Phys. Rev. Lett. 84
Methods Phys. Res. B 245 337 5090
McNeur J, Kozak M, Ehberger D, Schönenberger N, Tafel A, Popov V 1999 JETP Lett. 70 502
Li A and Hommelhoff P 2016 J. Phys. B: At. Mol. Opt. Phys. Porras M A 2009 Opt. Lett. 34 1546
49 034006 Priebe K E, Rathje C, Yalunin S V, Hohage T, Feist A,
McPherson A, Gibson G, Jara H, Johann U, Luk T S, McIntyre I A, Schäfer S and Ropers C 2017 Nat. Photonics 11 793
Boyer K and Rhodes C K 1987 J. Opt. Soc. Am. B 4 595 Putnam W P, Hobbs R G, Keathley P D, Berggren K K and
Miller M K, Cerezo A, Hetherington M G and Smith G D W 1996 Kärtner F X 2017 Nat. Phys. 13 335
Atom Probe Field Ion Microscopy (Oxford: Oxford University Quinonez E, Handali J and Barwick B 2013 Rev. Sci. Instrum. 84
Press) 103710
Milošević D B, Paulus G G, Bauer D and Becker W 2006 J. Phys. B: Rácz P, Ayadi V and Dombi P 2018 J. Opt. 20 015501
At. Mol. Opt. Phys. 39 R203 Rácz P and Dombi P 2011 Phys. Rev. A 84 063844
Morimoto Y and Baum P 2018 Nat. Phys. 14 252 Rácz P, Irvine S E, Lenner M, Mitrofanov A, Baltuška A,
Müller E W 1936 Z. Tech. Phys. 37 838 Elezzabi A Y and Dombi P 2011 Appl. Phys. Lett. 98 111116
Müller E W 1965 Science 149 591 Rácz P, Pápa Z, Márton I, Budai J, Wróbel P, Stefaniuk T, Prietl C,
Muller H G, Tip A and van der Wiel M J 1983 J. Phys. B: At. Mol. Krenn J R and Dombi P 2017 Nano Lett. 17 1181
Opt. Phys. 16 L679 Radoń T 1998 Prog. Surf. Sci. 59 331 Proc. 19th International
Müller M, Kravtsov V, Paarmann A, Raschke M B and Ernstorfer R Seminar Surface Phys. Meclewski Symp.
2016 ACS Photonics 3 611 Rathje T, Johnson N G, Möller M, Süßmann F, Adolph D, Kübel M,
Müller M, Paarmann A and Ernstorfer R 2015 Nat. Commun. 5 5292 Kienberger R, Kling M F, Paulus G G and Sayler A M 2012
Mulser P, Uryupin S, Sauerbrey R and Wellegehausen B 1993 Phys. J. Phys. B: At. Mol. Opt. Phys. 45 074003
Rev. A 48 4547 Reed B, Armstrong M, Browning N, Campbell G, Evans J,
Mustonen A, Beaud P, Kirk E, Feurer T and Tsujino S 2011 Appl. LaGrange T and Masiel D 2009 Microsc. Microanal. 15 272
Phys. Lett. 99 103504 Reiss H R 1980 Phys. Rev. A 22 1786

31
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Rethfeld B, Kaiser A, Vicanek M and Simon G 2002 Phys. Rev. B Taflove A and Hagness S 2005 Computational Electrodynamics: the
65 214303 Finite-Difference Time-Domain Method, Artech House
Riffe D M, Wang X Y, Downer M C, Fisher D L, Tajima T, Antennas and Propagation Library (Boston: Artech House)
Erskine J L and More R M 1993 J. Opt. Soc. Am. B 10 1424 Thomas S, Holzwarth R and Hommelhoff P 2012 Opt. Express 20
Ropers C, Elsaesser T, Cerullo G, Zavelani-Rossi M and Lienau C 13663
2007a New J. Phys. 9 397 Thomas S, Krüger M, Förster M, Schenk M and Hommelhoff P
Ropers C, Neacsu C C, Elsaesser T, Albrecht M, Raschke M B and 2013 Nano Lett. 13 4790
Lienau C 2007b Nano Lett. 7 2784 Thomas S, Wachter G, Lemell C, Burgdörfer J and Hommelhoff P
Ropers C, Solli D R, Schulz C P, Lienau C and Elsaesser T 2007c 2015 New J. Phys. 17 063010
Phys. Rev. Lett. 98 043907 Tóth C, Farkas G and Vodopyanov K L 1991 Appl. Phys. B 53 221
Runge E and Gross E K U 1984 Phys. Rev. Lett. 52 997 Tsong T T 1990 Atom-Probe Field Ion Microscopy (Cambridge:
Rybka T, Ludwig M, Schmalz M F, Knittel V, Brida D and Cambridge University Press)
Leitenstorfer A 2016 Nat. Photonics 10 667 Tsujino S, Beaud P, Kirk E, Vogel T, Sehr H, Gobrecht J and
Salvat F, Jablonski A and Powell C J 2005 Comput. Phys. Commun. Wrulich A 2008 Appl. Phys. Lett. 92 193501
165 157 Tsujino S, le Pimpec F, Raabe J, Buess M, Dehler M, Kirk E,
Sansone G et al 2006 Science 314 443 Gobrecht J and Wrulich A 2009 Appl. Phys. Lett. 94 093508
Sarid D and Challener W 2010 Modern Introduction to Surface Udem T, Holzwarth R and Hänsch T W 2002 Nature 416 233
Plasmons (Cambridge: Cambridge University Press) Udem T, Reichert J, Holzwarth R and Hänsch T W 1999 Phys. Rev.
Savage K J, Hawkeye M M, Esteban R, Borisov A G, Lett. 82 3568
Aizpurua J and Baumberg J J 2012 Nature 491 574 Uiberacker M et al 2007 Nature 446 627
Schafer K J, Yang B, DiMauro L F and Kulander K C 1993 Phys. Vampa G and Brabec T 2017 J. Phys. B: At. Mol. Opt. Phys. 50
Rev. Lett. 70 1599 083001
Schenk M, Krüger M and Hommelhoff P 2010 Phys. Rev. Lett. 105 Vampa G, McDonald C R, Orlando G, Klug D D, Corkum P B and
257601 Brabec T 2014 Phys. Rev. Lett. 113 073901
Schenk M, Krüger M and Hommelhoff P 2011 Frequency Control Vasa P, Ropers C, Pomraenke R and Lienau C 2009 Laser Photonics
and the European Frequency and Time Forum (FCS), 2011 Rev. 3 483
Joint Conf. IEEE Int. 1–3 Vesseur E J R, de Waele R, Kuttge M and Polman A 2007 Nano
Schertz F, Schmelzeisen M, Kreiter M, Elmers H-J and Lett. 7 2843
Schönhense G 2012 Phys. Rev. Lett. 108 237602 Vilayurganapathy S et al 2013 Appl. Phys. Lett. 103 161112
Scholl J A, García-Etxarri A, Koh A L and Dionne J A 2013 Nano Vogelgesang S, Storeck G, Horstmann J G, Diekmann T, Sivis M,
Lett. 13 564 Schramm S, Rossnagel K, Schäfer S and Ropers C 2018 Nat.
Schottky W 1914 Physik. Z. 15 872 Phys. 14 184
Schötz J et al 2018 Phys. Rev. A 97 013413 Vogelsang J, Robin J, Nagy B J, Dombi P, Rosenkranz D, Schiek M,
Schröder B, Sivis M, Bormann R, Schäfer S and Ropers C 2015a Groß P and Lienau C 2015 Nano Lett. 15 4685
Appl. Phys. Lett. 107 231105 Volkov D M 1935 Z. Phys. 94 250
Schröder B et al 2015b Phys. Rev. B 92 085411 Vurpillot F, Gault B, Vella A, Bouet M and Deconihout B 2006
Seiffert L, Henning P, Rupp P, Zherebtsov S, Hommelhoff P, Appl. Phys. Lett. 88 094105
Kling M F and Fennel T 2017a J. Mod. Opt. 64 1096 Vurpillot F, Houard J, Vella A and Deconihout B 2009 J. Phys. D:
Seiffert L et al 2017b Nat. Phys. 13 766 Appl. Phys. 42 125502
Seiffert L, Paschen T, Hommelhoff P and Fennel T 2018 J. Phys. B: Wachter G 2014 Simulation of condensed matter dynamics in strong
At. Mol. Opt. Phys. 51 134001 femtosecond laser fields PhD Thesis Vienna University of
Sivis M, Pazos-Perez N, Yu R, Alvarez- Puebla R, Technology http://repositum.tuwien.ac.at/urn:nbn:at:at-
García de Abajo F J and Ropers C 2018 Communications ubtuw:1-65500
Physics Wachter G, Lemell C, Burgdörfer J, Schenk M, Krüger M and
Smirnova O and Ivanov M 2014 Multielectron high harmonic Hommelhoff P 2012 Phys. Rev. B 86 035402
generation: simple man on a complex plane Attosecond and Wang F, Liu W, He L, Li L, Wang B, Zhu X, Lan P and Lu P 2017
XUV Physics (Weinheim: Wiley) pp 201–56 Phys. Rev. A 96 033407
Sonnenberg H, Heffner H and Spicer W 1964 Appl. Phys. Lett. 5 95 Wendelen W, Mueller B Y, Autrique D, Bogaerts A and Rethfeld B
Sönnichsen C, Franzl T, Wilk T, von Plessen G, Feldmann J, 2013 Appl. Phys. Lett. 103 221603
Wilson O and Mulvaney P 2002 Phys. Rev. Lett. 88 077402 Wessel J 1985 J. Opt. Soc. Am. B 2 1538
Spence J 2013 High-Resolution Electron Microscopy (Oxford: Wimmer L, Herink G, Solli D R, Yalunin S V,
Oxford University Press) Echternkamp K E and Ropers C 2014 Nat. Phys. 10 432
Stöckle R M, Suh Y D, Deckert V and Zenobi R 2000 Chem. Phys. Wimmer L, Karnbach O, Herink G and Ropers C 2017 Phys. Rev. B
Lett. 318 131 95 165416
Stockman M I 2004 Phys. Rev. Lett. 93 137404 Wittmann T, Horvath B, Helml W, Schatzel M G, Gu X,
Stockman M I and Hewageegana P 2007 Appl. Phys. A 89 247 Cavalieri A L, Paulus G G and Kienberger R 2009 Nat. Phys.
Stockman M I, Kling M F, Kleineberg U and Krausz F 2007 Nat. 5 357
Photonics 1 539 Wolf D, Schumacher T and Lippitz M 2016 Nat. Commun. 7 10361
Storeck G, Vogelgesang S, Sivis M, Schäfer S and Ropers C 2017 Yalunin S V, Gulde M and Ropers C 2011 Phys. Rev. B 84 195426
Struct. Dyn. 4 044024 Yalunin S V, Herink G, Solli D R, Krüger M, Hommelhoff P,
Sun Q, Ueno K, Yu H, Kubo A, Matsuo Y and Misawa H 2013 Diehn M, Munk A and Ropers C 2013 Ann. Phys. 525 L12
Light Sci. Appl. 2 e118 Yalunin S V, Schröder B and Ropers C 2016 Phys. Rev. B 93
Süßmann F and Kling M F 2011 Phys. Rev. B 84 121406(R) 115408
Süßmann F et al 2015 Nat. Commun. 6 7944 Yanagisawa H 2013 Ann. Phys. (Berlin) 525 126
Swanwick M E, Keathley P D, Fallahi A, Krogen P R, Laurent G, Yanagisawa H, Ciappina M, Hafner C, Schötz J, Osterwalder J and
Moses J, Kärtner F X and Velásquez-García L F 2014 Nano Kling M F 2017 Sci. Rep. 7 12661
Lett. 14 5035 Yanagisawa H, Hafner C, Doná P, Klöckner M, Leuenberger D,
Szriftgiser P, Guéry-Odelin D, Arndt M and Dalibard J 1996 Phys. Greber T, Hengsberger M and Osterwalder J 2009 Phys. Rev.
Rev. Lett. 77 4 Lett. 103 257603

32
J. Phys. B: At. Mol. Opt. Phys. 51 (2018) 172001 Topical Review

Yanagisawa H, Hafner C, Doná P, Klöckner M, Leuenberger D, Yavuz I, Bleda E A, Altun Z and Topcu T 2012 Phys. Rev. A 85 013416
Greber T, Osterwalder J and Hengsberger M 2010 Phys. Rev. Yuan H, He L, Wang F, Wang B, Zhu X, Lan P and Lu P 2018 Opt.
B 81 115429 Lett. 43 931
Yanagisawa H, Hengsberger M, Leuenberger D, Klöckner M, Yudin G L and Ivanov M Y 2001 Phys. Rev. A 64 013409
Hafner C, Greber T and Osterwalder J 2011 Phys. Rev. Lett. Zhang P and Lau Y Y 2016 Sci. Rep. 6 19894
107 087601 Zherebtsov S et al 2011 Nat. Phys. 7 656
Yanagisawa H, Schnepp S, Hafner C, Hengsberger M, Kim D E, Zherebtsov S et al 2012 New J. Phys. 14 075010
Kling M F, Landsman A, Gallmann L and Osterwalder J 2016 Zhu W, Esteban R, Borisov A G, Baumberg J J, Nordlander P,
Sci. Rep. 6 35877 Lezec H J, Aizpurua J and Crozier K B 2016 Nat. Commun. 7
Yang D-S, Mohammed O F and Zewail A H 2010 Proc. Natl Acad. 11495
Sci. USA 107 14993 Zuloaga J, Prodan E and Nordlander P 2010 ACS Nano 4 5269

33

You might also like